symbolic design and optimization techniques for analog ...kp384vk8999/thesis_sidseth... · symbolic...

147
SYMBOLIC DESIGN AND OPTIMIZATION TECHNIQUES FOR ANALOG INTEGRATED CIRCUITS A DISSERTATION SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Siddharth Seth June 2013

Upload: ngodung

Post on 05-Jul-2018

227 views

Category:

Documents


0 download

TRANSCRIPT

SYMBOLIC DESIGN AND OPTIMIZATION TECHNIQUES FOR ANALOG

INTEGRATED CIRCUITS

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Siddharth Seth

June 2013

http://creativecommons.org/licenses/by-nc-sa/3.0/us/

This dissertation is online at: http://purl.stanford.edu/kp384vk8999

© 2013 by Siddharth Seth. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial-Share Alike 3.0 United States License.

ii

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Boris Murmann, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Mark Horowitz

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Bruce Wooley

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

iv

v

Abstract

An analog circuit design problem typically has many acceptable solutions.

However, within the very broad design space, there will usually exist one optimal

design that minimizes (or maximizes) one of the objectives, given a constraint on the

other metrics. The rising complexity of the circuits and the absence of closed-form

expressions for certain metrics (like total integrated noise) have led to a SPICE-

simulation-based numerical approach to analog circuit design and optimization, which

is very slow for circuits comprising more than a handful of transistors. The research

presented in this dissertation focuses on symbolic design and optimization techniques

for analog integrated circuits. These techniques are based on computer optimization

programs that use closed-form symbolic expressions for all relevant performance

metrics of the analog circuit, bypassing the need to interface with a circuit simulator.

In the first part of this work, we deal with the problem of computing total

integrated noise in an analog circuit. We demonstrate a technique to compute the total

integrated noise by visual inspection in linear, passive networks, and then extend the

technique to show how one can symbolically integrate a general noise transfer function

of any order to get closed-form expressions for total integrated noise. Such expressions

were not readily available and had prevented the adoption of symbolic analysis in the

design and optimization of noise limited analog circuits. Compared to previously known

methods, this technique is efficient in terms of computation cycles and memory

requirement, and provides the answer in a single step.

We next present three proof-of-concept examples that illustrate how symbolic

analysis can be applied to the design and optimization of representative analog blocks.

The presented techniques are general, and taken together, can help provide a circuit

designer with the best design, find sensitivities to circuit parameters, and enable rapid

design portability to different sets of specification or process corners.

In the first example, we present a nested-Miller-compensated three-stage

operational transconductance amplifier for use in high-speed switched-capacitor circuits.

Simulation results show that the 90-nm prototype amplifier achieves a 0.1 % dynamic

error settling time of 2.53 ns with a total integrated noise of 240 μVrms, while

consuming 5.2 mW from a 1-V power supply.

vi

In the second example, we present the design and optimization of continuous-

time active-RC and gm-C low-pass filters. Starting from a given LC ladder-filter

realization, we develop a systematic method of choosing the right optimization

variables and using signal-flow-graph manipulations to convert a given LC ladder-filter

realization into the final analog circuit. This is done in such a way that the symbolic

expressions for noise, power and area turn out to be posynomial functions, enabling the

formulation of the design and optimization problem as a geometric program (GP) that

can be quickly solved to get the globally-optimal solution. One of the limitations in

such filters is the problem of device mismatch and variability. As a solution, critical

components like transconductors, resistors and capacitors are usually chosen to be

integer multiples of each other. We add such practical constraints to the optimization

problem, and branch-and-bound techniques are used to solve the resulting mixed-

integer GP (MIGP).

Finally, in the third example, we present the analysis, design, and measurement

results of a low-noise, low-power, series-resonant MEMS oscillator at 20 MHz that

consists of a high-Q differential resonator, wire-bonded to a high-gain CMOS

transimpedance amplifier (TIA). Symbolic analysis is used to evaluate the impact of

TIA bandwidth on the oscillator frequency and phase noise, and accordingly a suitable

topology is chosen and optimized. Measurement results show that the designed

oscillator compares favorably to the state-of-the-art in terms of its circuit design figure-

of-merit.

vii

Acknowledgments

When I joined Stanford for the M.S. / Ph.D. program in 2006, one of my favorite

activities here was visiting the ‘Theses’ section of the Terman Engineering Library,

picking up a random thesis, feeling the weight of the red, hard-bound thesis books, and

going through the many passionately-written acknowledgments in them. Back then, it

was the only section in those theses I could understand. Nonetheless, being

impractically imaginative, I had started writing my own thesis acknowledgment in my

mind, even before I cleared the Ph.D. qualifying examination, or had an advisor, or had

figured out a research topic.

What followed those initial years has been a whirlwind journey with deep lows

in the middle, but immensely gratifying highs at the end. As I write these last

paragraphs of my thesis, I remember the many times (many, many times) in which I had

doubted whether I will be able to come this far. For making this journey possible, I

would wholeheartedly like to thank the people below.

… At a place like Stanford, it is easy to get overwhelmed by the sheer

excellence of the many accomplished faculty. However, the single biggest contribution

to my 'coming of age' at Stanford, has been from a genuinely excellent person who is a

fine teacher and a wonderful adviser ‐ Prof. Boris Murmann. I had completely jumbled

up my research plans, and had nearly decided to quit the Ph.D. program in mid-2010. I

would not be graduating today, had it not been for Prof. Murmann who took me in as

his student at that time. I am yet to meet or even hear about anyone who works harder

than him. Thank you Prof. Murmann for teaching me circuit design, for tonnes of

motivation, for several ‘aa-haa’ moments, and for always being accessible. Thank you

for patiently editing all my technical write-ups and for teaching me the intellectual merit

of perfectionism. Above all, thank you for, by example, making me love my work again.

Vielen Dank, Professor Murmann.

… Thank you Prof. Bruce Wooley and Prof. Mark Horowitz for being on my

orals and reading committees, and for taking time out during the long memorial-day

weekend to read my thesis draft in such detail. To see them being so generous to us

students, in spite of their super-star statures, is a great learning experience.

viii

… Thank you Prof. Amin Arbabian for being on my orals committee and Prof.

Thomas Kenny for chairing the orals committee. Thank you both for the enlightening

discussions I had with you about my research.

… Thank you Prof. Donald Cox for the opportunity to TA EE344 and for letting

me ride and drive your Tesla Roadster on 280 - one of the most memorable experiences

at Stanford!

... Thank you Prof. Thomas H. Lee for the opportunity to TA for EE314 and

EE414 four times, and for all that I learned from you.

.. Thank you Ann Guerra for your charm and for wielding real power in getting

things done, Joe Little for being available even on weekends and midnights to help me

with server issues, Keith Gaul for always fulfilling my supply- and equipment-related

requests with a smile, Pauline Prather for getting us the last mile access to our chips

with impeccable wire-bonding jobs and for never saying no, no matter how late we

made her work at nights. You all are, truly, the super-women and super-men of our

department.

… Thank you Robert Bosch Stanford Graduate Fellowship (SGF) for financial

support during my initial years at Stanford, and the Stanford Rethinking Analog Design

(RAD) for funding my research work. I would also like to thank Invensense, Inc. and

Mike Daneman for chip fabrication.

… My mentors in the industry, Shahriar Rabii, Ron Ho and Bruno Garlepp,

thank you for patiently answering my career-related queries and for sharing your own

stories. Thank you Thomas Cho and Chih-Wei Yao at Samsung Research, Hae-Chang

Lee and Vinod Menon at SiTime, David Su at Qualcomm Atheros, Ari Vauhkonen at

Maxim, and Changku Hwang at Oracle, for making the job interview process an

exciting and humbling experience for me.

… My first managers, Alec Woo and Feng Zhao at Microsoft Research - thank

you for believing in me and for recommending my application at Stanford.

… Thank you Gadgil Sir for singling me out of the 300 bright students in your

class in high school and for never stopping to believe in my capabilities. I would have

never qualified for IIT or Stanford, had it not been for you. Thank you for teaching me

calculus and for making me appreciate the beauty of it all. Thank you Mr. Shaikh and

ix

Sahastrabuddhe Ma’am for being the first teachers to teach me circuit design. Thank

you Prof. N. B. Chakrabarti (IIT Kharagpur) for all that I learned from you.

… Thank you Rabin Patra and Akshat Verma for being my childhood heroes

and for inspiring me to get through IIT and grad-school.

… Thank you Drew Hall for being such an inspiration - you are one of the most

hard-working students I have ever met. Thank you Maryam for managing to remain

incredibly smart and incredibly humble at the same time. Thank you Pedram for

teaching me so many things and for our many chats over email. I would also like to

thank the past and present members of the Murmann and Wooley group for their useful

discussions and feedback. In particular, I will fondly cherish the times spent with Noam,

Mohammad, Ross, Alireza, Kamal, Vaibhav, Martin, Alex and Shasha.

… Thank you Isaac Martinez-Garcia for writing one of the most inspiring and

heartfelt acknowledgments I have read, something that kept me motivated when the

going got tough.

… Thank you Amit Trivedi and A R Rahman for creating foot-tapping and

melodious Hindi film music, Amitabh Bhattacharya for beautiful poetry, and Roger

Ebert for making movies (and life) even more special with his charming commentary.

Thank you for brightening up many of my gloomy days with your music and words.

… Thank you Chintan and Vijayeta, Anshuman and Divya, Gaurav Thareja,

Sakshi and Ratndeep, Niket and Ruchika, Vinay and Vandana, Dhiren, Amit, and Rahul.

Your friendship has meant the world to me. I am blessed.

… Thank you, Chintan. Thank you for always making me feel like a rockstar.

Thank you for that particular email, for your advice and your words of encouragement.

Proud of you!

… My family-in-law: Mummy, Papa, Chhavi, Cheenu, Mona, Santhosh, Deepa

Masi, Neeraj Mausaji, Nandan, Vrinda - I cannot thank you enough for the love you

have showered. Thank you!

… My family in the bay area: Munni Masi, Neha, Mausaji, Bade Nana-Nani,

Chhote Nana-Nani, Sonu Masi, Sam, Matthew, Pinky Masi, Jay, Dev and Shay - thank

you for taking such good care of us. Munni Masi, you are the most self-less person I

know - thank you for showering me with your motherly love.

x

… I am what I am because of my parents and sister. Maa-Baba and Shalu - I

have missed you a lot all these years. I know that you are my biggest fans and my

loudest cheerleaders. The smallest of my achievements make you go gaga. Thank you

Maa-Baba for making sure Shalu and I got the best education possible, even if it meant

that Baba stay alone in remote towns. Baba, thank you for all the sacrifices you have

made and for never complaining. Thank you for teaching me sine-waves and for sharing

EE jokes from your own college days. Maa, while I do have many differences of

opinion with you, I admire your spirit and carefree attitude. Thank you for pampering

me and Shalu so much. I must confess it has spoiled me a bit. Shalu, you seem to have

suddenly grown up, and I rue the fact that I was not around to see you all these years.

How I wish we could relive our wonderful childhood years.

… Kanupriya, thank you for being my best friend and my most ardent admirer. I

am very proud of the work you have accomplished in your own graduate work while

putting up with my tantrums and all the additional baggage that comes with being my

full-time caretaker. I sometimes still cannot believe how someone like you could agree

to be with someone like me. I will always cherish our story that began with the early

morning walks in Kharagpur and continued with the late night dinners amidst our chips

at the laboratory at Stanford. As someone once said, there is no limit to the possibilities

that we two have together. I am sorry for not being there in the many ways I should

have. Thank you for being a witness to my life.

For all the hard work and patience that has gone in making me what I am, I

dedicate this thesis to Maa, Baba and Kanupriya. You have all loved me far more than I

can ever reciprocate.

To the reader who stuck around until the end and read this very long and

informal emotional outpouring in an otherwise formal dissertation, I would like to

extend a heartfelt thank you for your patience. May you stand on the shoulders of giants,

like I did, and discover both: the world of EE, and yourself. Maybe one of these readers

is me, reading this a few or tens of years from the day all this got written. Remember, I

was never alone.

xi

Table of Contents

Abstract......... ................................................................................................................ v

Acknowledgments ....................................................................................................... vii

List of Tables .............................................................................................................. xv

List of Figures ........................................................................................................... xvii

List of Symbols and Acronyms................................................................................. xxi

CHAPTER 1 Introduction ........................................................................................... 1

1.1 Research Motivation................................................................................................. 1

1.1.1 Analog Circuit Design Space ................................................................ 1

1.1.2 Analog Circuit Analysis Methods ......................................................... 3

1.1.3 Running an Analog Optimization Loop ................................................ 4

1.1.4 Prior Work in Analog Circuit Design and Optimization ...................... 6

1.2 Organization of Thesis Dissertation ......................................................................... 7

CHAPTER 2 Symbolic Computation of Total Integrated Noise ............................. 9

2.1 Background .............................................................................................................. 9

2.2 Previously Known Methods ................................................................................... 10

2.2.1 Residue Theorem ................................................................................ 11

2.2.2 Lyapunov Equations ........................................................................... 12

2.3 Computing Total Integrated Noise by Inspection in Linear, Passive Networks .... 13

2.4 Track Mode Noise in Switched Capacitor Circuits ................................................ 18

2.5 Exact Closed-Form Expressions for Total Integrated Noise in Arbitrary Circuits …

............................................................................................................ 21

2.6 Conclusion ............................................................................................................ 27

CHAPTER 3 Settling Time and Noise Optimization of a Three-Stage OTA ....... 29

3.1 Motivation ............................................................................................................ 29

3.2 Design and Optimization Framework .................................................................... 31

3.2.1 Model Description .............................................................................. 32

3.2.2 Equations and Constraints Formation ................................................. 34

3.2.3 Running the Nonlinear Constrained Optimization Program ............... 38

3.3 Final Design ........................................................................................................... 40

xii

3.4 Conclusion ............................................................................................................. 42

CHAPTER 4 Design and Optimization of Continuous-Time, gm-C and Active-RC

Filters ................................................................................................... 43

4.1 Motivation ............................................................................................................. 43

4.1.1 Prior Work in Filter Design and Optimization .................................... 46

4.2 Geometric Program (GP) and Mixed-Integer Geometric Program (MIGP) ......... 47

4.3 Design and Optimization Framework ..................................................................... 49

4.3.1 Elliptic Filter Structure ........................................................................ 49

4.3.2 Systematic Signal-Flow-Graph Based Modeling ................................ 51

4.4 Design and Optimization of the gm-C Filter Topology ........................................... 57

4.4.1 Transconductor Structure .................................................................... 57

4.4.2 Equations and Constraints Formation ................................................. 58

4.4.3 GP for Optimization of gm-C Filter Design ......................................... 61

4.4.4 MIGP for Dynamic Range Optimization ............................................ 67

4.5 Design and Optimization of active-RC Filter Topology ........................................ 71

4.5.1 Op-Amp Structure ............................................................................... 71

4.5.2 Equations and Constraints Formation ................................................. 71

4.5.3 Geometric Program for Noise Minimization ...................................... 76

4.6 Conclusion ............................................................................................................. 78

CHAPTER 5 Design and Optimization of a Gain-Tunable TIA for a MEMS

Oscillator .............................................................................................. 81

5.1 Background ............................................................................................................. 81

5.2 Breathe-Mode Ring Resonator ............................................................................... 83

5.3 TIA Based Series Resonant Oscillators .................................................................. 85

5.3.1 Symbolic Analysis of Oscillation Frequency, Required TIA Gain and

Phase Noise ......................................................................................... 86

5.3.2 TIA Bandwidth .................................................................................... 88

5.3.3 Symbolic Analysis Summary .............................................................. 89

5.4 TIA Design and Optimization ................................................................................ 90

5.4.1 TIA Topology ...................................................................................... 90

xiii

5.4.2 Design and Optimization .................................................................... 93

5.4.3 Transistor Level Implementation ........................................................ 94

5.5 Measurement Results ............................................................................................. 98

5.6 Performance Summary and Comparison .............................................................. 100

5.6.1 Circuit Design Figure-of-Merit in Series-Resonant Oscillators ....... 100

5.6.2 Performance Comparison ................................................................. 102

5.7 Conclusion .......................................................................................................... 102

CHAPTER 6 Conclusions and Future Outlook..................................................... 105

6.1 Summary .......................................................................................................... 105

6.2 Areas of Future Work ........................................................................................... 106

APPENDIX A Proof of the Bode Theorem for Impedance/Admittance Integrals

............................................................................................................ 109

APPENDIX B Solution of Equation (2-30) ........................................................... 113

Bibliography….......................................................................................................... 117

xiv

xv

List of Tables

Table 2-1. Closed-form symbolic expressions for . ...................................... 25

Table 3-1. Design specifications for SC gain stage........................................................ 31

Table 3-2. design choices. ................................................................................. 33

Table 3-3. Design parameters. ........................................................................................ 40

Table 3-4. Performance summary. ................................................................................. 41

Table 3-5. Comparison between calculated and simulated total integrated thermal

voltage noise. ................................................................................................ 42

Table 4-1. Filter design specifications. .......................................................................... 49

Table 5-1. Ring resonator dimensions. ........................................................................... 83

Table 5-2. Symbolic analysis summary for series resonant oscillators. ......................... 90

Table 5-3. TIA design parameters. ................................................................................. 93

Table 5-4. TIA power consumption breakdown. ........................................................... 99

Table 5-5. MEMS oscillator performance summary and comparison. ........................ 102

xvi

xvii

List of Figures

Figure 1-1. Analog circuit design space — one problem, many solutions. ..................... 2

Figure 1-2. (a) Signal and noise in an electronic system. (b) Analog circuit design trade-

offs. (c) The analog circuit design problem formulated as a constrained

optimization program. .................................................................................... 3

Figure 1-3. Analog design and optimization loop. ........................................................... 5

Figure 2-1. Computation of total integrated voltage noise by inspection. ..................... 16

Figure 2-2. Computation of total integrated current noise by inspection. ...................... 17

Figure 2-3. Charge-redistribution track-and-hold amplifier stage. ................................ 18

Figure 2-4. Circuit model for noise analysis during track phase.................................... 18

Figure 2-5. Computing mean squared noise voltages by inspection. ............................. 20

Figure 2-6. Comparison of time taken by three methods of symbolic integration. ........ 27

Figure 3-1. Model of an SC gain stage during charge redistribution. ............................ 29

Figure 3-2. Transistor intrinsic gain across technology nodes. ............................. 30

Figure 3-3. Transistor-level implementation of the OTA (bias network not shown). All

NMOS bodies are tied to ground, all PMOS bodies are tied to ............ 32

Figure 3-4. Behavioral model of nested-Miller-compensated OTA in an SC amplifier..

...................................................................................................................... 34

Figure 3-5. Normalized step-response. ........................................................................... 35

Figure 3-6. Running the optimization program. ............................................................. 39

Figure 3-7. Settling to within dynamic error. .................................................... 41

Figure 3-8. Comparison between calculated and simulated noise PSDs. ...................... 41

Figure 4-1. Analog filter synthesis steps. ....................................................................... 45

Figure 4-2. LC ladder prototype and passive component values. .................................. 50

Figure 4-3. Single-ended circuit for gm-C topology. ...................................................... 50

Figure 4-4. Single-ended circuit for active-RC topology. .............................................. 51

Figure 4-5. Differential circuits for (a) gm-C topology, (b) active-RC topology ........... 52

Figure 4-6. Signal-flow-graph of circuit in Figure 4-2. ................................................. 53

Figure 4-7. Modified SFG of circuit in Figure 4-2......................................................... 53

Figure 4-8. Modified SFG with internal node-voltage scaling. ..................................... 54

xviii

Figure 4-9. Final, fully-scaled SFG of circuit in Figure 4-2. .......................................... 54

Figure 4-10. Relating the SFG to circuit elements for gm-C and active-RC topologies. 55

Figure 4-11. Transconductor stage. ................................................................................ 57

Figure 4-12. Gain to internal nodes in the gm-C topology before node-voltage scaling. 61

Figure 4-13. Gain to internal nodes in the gm-C topology after node voltage scaling. ... 62

Figure 4-14. Pareto-optimal curve for the 5th

order elliptic gm-C filter. ......................... 64

Figure 4-15. Gain to internal nodes in the gm-C topology for .................. 66

Figure 4-16. Relative change in total integrated noise and dynamic range in the gm-C

topology as a function of . ....................................................................... 67

Figure 4-17. Gains to the internal nodes in the gm-C topology, all transconductors equal.

.................................................................................................................... 68

Figure 4-18. Gains to internal nodes in the gm-C topology when all transconductors are

integral multiples of a unit transconductor, ....................... 70

Figure 4-19. Dynamic range as a function of in the gm-C topology. ............. 70

Figure 4-20. Op-amp structure. ...................................................................................... 71

Figure 4-21. Gains to the internal nodes in the active-RC topology. ............................. 77

Figure 4-22. Relative change in total integrated noise and dynamic range as a function

of in the active-RC topology. ................................................................. 77

Figure 4-23. Change in noise contributions from the resistors and from the op-amps in

the active-RC topology as a function of ........................................... 78

Figure 5-1. Timing reference in a system-on-chip. ........................................................ 81

Figure 5-2. WLAN system-in-package module. ............................................................. 82

Figure 5-3. Breathe-mode ring resonator. ....................................................................... 83

Figure 5-4. Electrical model for resonator near resonance. ............................................ 84

Figure 5-5. Series-resonant oscillator. ............................................................................ 85

Figure 5-6. Series-resonant oscillator model. ................................................................. 86

Figure 5-7. Phase noise in series-resonant oscillators. ................................................... 88

Figure 5-8. Maximally-flat versus peaking transfer function in the core TIA. .............. 89

Figure 5-9. Design and optimization strategy for TIA. .................................................. 90

Figure 5-10. Resistive-feedback TIA topology. ............................................................. 90

xix

Figure 5-11. TIA with tunable complex conjugate pole pair (biasing circuits not shown).

.................................................................................................................... 91

Figure 5-12. Circuit schematic of the TIA. .................................................................... 95

Figure 5-13. Circuit schematic of the OTA. ................................................................... 96

Figure 5-14. Generating the sense terminal bias. ........................................................... 97

Figure 5-15. ALC block diagram. .................................................................................. 97

Figure 5-16. System micrograph. ................................................................................... 98

Figure 5-17. Measurement of TIA gain.......................................................................... 98

Figure 5-18. Open-loop measurement of the MEMS resonator characteristics. ............ 99

Figure 5-19. Closed-loop measurements. ....................................................................... 99

Figure 5-20. Phase noise in a PLL. .............................................................................. 101

xx

xxi

List of Acronyms

ADC - Analog-to-Digital Converter

AGC - Automatic Gain Control

ALC - Automatic Level Control

BAW - Bulk Acoustic Wave

CMFB - Common Mode Feed-Back

CMOS - Complementary Metal-Oxide Semiconductor

CPU - Central Processing Unit

CT - Continuous Time

DR - Dynamic Range

FoM - Figure of Merit

GP - Geometric Program

IC - Integrated Circuit

IP - Intellectual Property

KCL - Kirchhoff’s Current Law

KVL - Kirchhoff’s Voltage Law

LHP - Left Hand Plane

LTI - Linear Time Invariant

LTV - Linear Time Variant

MEMS - Micro Electro Mechanical System

MIGP - Mixed-Integer Geometric Program

NTF - Noise Transfer Function

OTA - Operational Transconductance Amplifier

PSD - Power Spectral Density

RF - Radio Frequency

RHP - Right Hand Plane

SAW - Surface Acoustic Wave

SC - Switched Capacitor

SFG - Signal Flow Graph

SNR - Signal to Noise Ratio

SoC - System on Chip

xxii

THA - Track and Hold Amplifier

TIA - Trans-Impedance Amplifier

WLAN - Wireless Local Area Network

1

CHAPTER 1

Introduction

1.1 Research Motivation

We live in an analog world. Although we usually condition and convert the

analog information into digital information for processing or storage into digital media,

we essentially create and consume information in the analog domain. The number of

digital components in modern-day portable electronic devices has increased because of

the emergence of advanced digital signal processing techniques and digital modulation

schemes. However, these devices have also evolved from having a single analog

interface, to now having multiple analog interfaces that process not only voice, but also

touch, motion, images, videos, etc. while supporting multiple wireless standards [1].

The analog circuits in these devices need to be high-performance, and consequently,

their design and optimization is very important. However, while digital design is based

on well-established practices supported by digital IP reuse, and automated synthesis

tools, analog design is based largely on designer intuition and experience and repetitive

simulation-based optimization. Consequently, although the analog part of a mixed-

signal IC is simpler in complexity than the digital part, it typically turns out to be the

bottleneck in getting a functional chip [2].

1.1.1 Analog Circuit Design Space

Analog circuits exist in a very broad design space. As an illustration, consider

the design of a low-pass filter. As shown in Figure 1-1, we can implement such a filter

using one of multiple topologies. Furthermore, once a topology is selected, we can

select multiple component values that implement the same end-to-end transfer function.

In effect, there exists a one-to-many relationship between an analog circuit design

problem and its solutions. One of these solutions has better performance (evaluated in

terms of performance metrics like noise, power consumption, area, etc.) compared to

the others, and that is the solution that should be implemented.

Chapter 1 - Introduction

2

Figure 1-1. Analog circuit design space — one problem, many solutions.

One of the most important metrics used to evaluate the performance of a given

analog circuit is its noise. In any electronic system, the signal we are trying to detect or

process exists with some background electronic noise, as shown in Figure 1-2 (a). The

ratio between the signal and the noise, i.e. the signal-to-noise ratio, determines the

fidelity of the system. As CMOS technology has scaled, the power supplies have gone

down, and that lowers the swing available for the signal. This makes noise performance

more critical. As shown in Figure 1-2 (b), the other performance metrics like power

consumption, bandwidth, gain, linearity, etc. trade off with noise and with each other. In

the light of such inter-woven relationships, the task of finding a good solution to an

analog design problem can be formulated into a constrained optimization program, as

shown in Figure 1-2 (c), where we want to optimize one of these performance metrics

subject to constraints on the other metrics.

5th Order Elliptic Low-Pass Filter

fc = 11 MHz

-1

-1

-1

-1

gm-C

Passive LC

Active-RC

Vin

Vout

Vin Vout

Vin Vout

Multiple topologies, multiple component values

Chapter 1 - Introduction

3

Figure 1-2. (a) Signal and noise in an electronic system. (b) Analog circuit design trade-offs. (c) The

analog circuit design problem formulated as a constrained optimization program.

1.1.2 Analog Circuit Analysis Methods

The first step in solving the analog circuit optimization problem is circuit

analysis. Once a topology is selected and the schematics have been created, the circuit

can be analyzed in one of two ways [3]:

Symbolic analysis: The circuit parameters, input frequency, input signal strength, etc.

are treated as symbols. Nodal analysis is done using KVL and KCL and

mathematical operations are performed on these symbols to formulate implicit or

explicit (i.e. closed-form) symbolic expressions for various performance metrics.

Numerical analysis: Numerical values are assigned to the circuit parameters, input

signal frequency, input signal strength and a computer-based circuit simulator like

SPICE [4] or Cadence Spectre [5] is used to perform a numerical computation of

the various performance metrics. The numerical values assigned to the circuit

components, etc. are varied and the performance metrics are numerically evaluated

at each point.

Over the past three decades, analog circuit design has evolved from a symbolic

analysis based approach to numerical analysis based approach. This has happened

because of the following reasons:

Rising complexity: Symbolic analysis can be performed by hand for circuits with

only a few transistors. Examples of such circuits are differential pair, single-stage or

Signal

Noise Noise

Power

Bandwidth

Linearity

Gain

optimize

subject to

one of the performance metrics

constraints on the other performance metrics

(a) (b)

(c)

Chapter 1 - Introduction

4

two-stage amplifiers, and filters with order less than two. However, modern-day

electronic devices require complex analog circuits for achieving the stringent design

specifications. Examples of such circuits are multi-stage amplifiers and high-order

continuous-time or switched-capacitor filters. Symbolic analysis by hand tends to be

tedious and error-prone for such circuits. While computer-based symbolic analyzers

like the symbolic math toolbox in MATLAB [6] are available for such a task, it is

not widely adopted by circuit designers.

Absence of closed-form expression: While it is straight-forward to find closed-form

symbolic expressions for metrics like bandwidth, gain, area, and power

consumption, methods for finding closed-form symbolic expressions for total

integrated noise are computationally expensive and lead to long computation times.

Such expressions are readily available only for noise integrals up to the fourth order

[7]. This has prevented the use of noise as an objective or constraint in symbolic

analysis based analog circuit design and optimization.

Transistor modeling: The mapping between a transistor’s dimensions and its

parameters like current , transconductance , intrinsic gain , etc. is very

complicated for deep-submicron CMOS technologies that use models like BSIM4

[8]. Consequently, for optimization methodologies that treat transistor dimensions

as the optimization variables, it is very difficult to formulate objective and

constraint equations. However, using the derived small-signal circuit parameters

like , , as the optimization variables along with the lookup table

methodology [9] to find the transistor dimensions, can help remove this bottleneck.

1.1.3 Running an Analog Optimization Loop

While using a computer-based simulator for numerical analysis of a circuit is

quick way to characterize a circuit, using it to design and optimize the circuit is a very

slow process. This is further illustrated by the decision diagram shown in Figure 1-3.

The designer starts with a certain set of specifications and a starting point. The

performance metrics are evaluated at the current design point. This can be done by

going back to the simulator, which is a very slow process, or by using a complete

symbolic model, which is faster. The optimization algorithm then changes the design

Chapter 1 - Introduction

5

point to make sure that the objective converges to the optimum and the constraints are

met. A variety of algorithms can be used for finding a feasible direction in which the

design should evolve [2]. Almost all such algorithms rely on computing the gradient of

the objective and constraint functions with respect to each design variable. Some

nonlinear programming algorithms also use the Hessian matrix (a square matrix of

second-order partial derivatives of the function) in order to ensure convergence and to

provide results that are more accurate. The gradients and Hessian matrices can be found

by two methods. The optimization algorithm can perturb each variable and use the

simulator to evaluate the objective and constraints and then compute the gradients and

Hessians using finite differences (a slow process), or simply use the symbolic model to

find the gradients directly by differentiation over the closed-form expressions of the

objective and constraints (a fast process). Sometimes, the optimizer can reach a design

point that is feasible, but finite differences around lead to an infeasible point,

causing the optimizer to diverge or halt prematurely. In such cases, providing gradients

and Hessians directly from closed-form expressions allows the optimizer to converge to

a solution [10], [11].

Figure 1-3. Analog design and optimization loop.

We thus conclude that finding the gradient using finite differences can be time

consuming, can lead to inaccurate results, and may even cause the optimizer to fail.

Specifications / Starting Point

Evaluate Objective, Constraints, 1st Order

Derivs., 2nd Order Derivs.

Simulator

Sym. Model

SPICE, Spectre, etc.

MATLAB, Maple, etc.

Change Design Variables using Optimization

Algorithm

noyes Design Converges?

Final Design Point

Quantify Behavior at Single Point in Design Space

Transient, Noise Simulations - Slow

Formula Based Calculation - FastQuantify Behavior at Infinite Points in Design Space

Chapter 1 - Introduction

6

This can be especially problematic when the optimizer needs to be run multiple times

with new starting points, in order to find the globally optimal design point. Each

optimizer run itself involves multiple transient, noise, and ac simulations. This leads to

impractically long run times if a circuit simulator is used, even for small and medium

sized circuits.

1.1.4 Prior Work in Analog Circuit Design and Optimization

A multitude of analog circuit design and optimization tools that interface with a

circuit simulator have been available for a long time. A few examples of such tools are

OASYS [12], IDAC [13], OPASYN [14], DELIGHT.SPICE [15], ASTRX/OBLX [16],

AMGIE [17], MAELSTROM [18], ANACONDA [19], ASF [20], and VASE [21]. A

brief review of such tools can be found in [2]. These tools differ from each other in the

way the optimizer interfaces with the circuit simulator. However, these tools utilize

little or no knowledge of the analog circuit available to the designer from intuition or

experience, and treat the circuit as just an abstracted mathematical problem.

Commercial tools like Virtuoso NeoCircuit [22] and the ‘digital-analog design’

methodology [23] use built-in parameterized modules and custom-built analog cells for

design and optimization. These are based on using techniques from digital circuit design

procedures, like function encapsulation, system validation, automatic electrical rules

checking, and IP reuse.

Symbolic analysis based optimization has been used for the design of robust

digital circuits [24]. In this approach, the symbolic model of the digital circuit lends the

associated optimization problem to be implemented as a geometric program, which is

then efficiently solved using interior point methods [10]. A similar approach had been

previously used for the design and optimization of a two-stage op-amp [25]. However,

in an attempt to make the final optimization problem as a geometric program, this

approach made simplifying assumptions like using spot noise instead of total integrated

noise, assuming a single pole system, etc. These simplifications work only for simple

circuits, and for such circuits, a simulator based optimization routine can provide as

good or better designs.

Chapter 1 - Introduction

7

1.2 Organization of Thesis Dissertation

The research presented in this thesis aims at developing design and optimization

methodologies for analog circuits that are:

Symbolic-analysis based, for fast and accurate implementation: The symbolic model,

developed using the symbolic toolbox in MATLAB [6] can be created such that it

closely matches the physical circuit. If all the circuit parameters including parasitic

and all the performance metrics can be expressed as closed-form expressions of the

design variables, there will be no need to interface with a circuit simulator.

Capable of dealing with complex circuits: The design procedures for circuits like

band-gap references, differential pairs, two-stage amplifiers, etc. have been well

known for a long time [26]. The developed methodologies should be able to deal

with complex circuits like multi-stage amplifiers and higher-order filters that lack a

quantitative design procedure because of the plethora of inter-related performance

metrics.

Generic and knowledge-based: The goal is to provide a set of techniques that are

general and can be suitably modified for application applied to any analog circuit

design problem, instead of providing a software CAD tool. The design intelligence

still comes from the designer, the techniques only aid in formalizing the designer’s

intent and creativity, while taking care of the laborious numerical part of the design

and optimization process. The methodologies should enable the designer to not only

find the best design, but also evaluate sensitivities, test rules of thumb and

conventional wisdom, and port the design to different sets of specifications and

process corners.

The remainder of the thesis is organized as follows: Chapter 2 deals with the

problem of computing closed-form expressions for total integrated noise. Next, Chapter

3, Chapter 4 and Chapter 5 respectively describe the design and optimization of a three-

stage OTA, continuous-time active filters and a CMOS TIA for MEMS applications.

Lastly, concluding remarks and future research directions are presented in Chapter 6.

Chapter 1 - Introduction

8

9

CHAPTER 2

Symbolic Computation of Total Integrated

Noise

2.1 Background

Noise analysis in an arbitrary, active or passive, linear, analog circuit proceeds

as follows. Assuming that all the noise sources are independent of each other, the noise

PSD at the output is found by summing the products of the individual source noise two-

sided PSD's , with the squares of the magnitudes of their corresponding noise

transfer functions (NTF) to the output, [27]. In the case of thermal noise, the

source PSD's are constant with frequency. Therefore, for the noise source, we can

write , a constant. We integrate the noise PSD at the output from zero to

infinity to get the mean square value of the noise [26],

(2-1)

where we have used the fact that is an even function of the frequency

A lot of analysis exists in the literature for the first part of this process, namely,

finding the PSD at the output. For a general, linear, time-invariant, amplifier based

circuit, noise PSD can be computed using adjoint theory [28], Padé approximation

based model reduction [29], etc. For analog filters [30], more detailed analyses are

available that make use of features specific to the particular filter topology or response

type. For example, noise performance has been analyzed for continuous time (CT)

active-RC and MOSFET-C filters [31]-[32], and operational-transconductance-

amplifier based gm-C filters [33]-[34]. Similar noise analysis in periodically switched

linear analog circuits like switched-capacitor (SC) filters is provided in [35] and [36].

Although SC filters are not time invariant, their noise analysis also involves integrating

Chapter 2 - Symbolic Computation of Total Integrated Noise

10

a noise transfer function (NTF) to obtain the total integrated noise, just as in CT filters

[36].

While the above-cited works provide expressions for the noise PSD, one is still

left with the task of integrating the PSDs to compute the noise in the relevant signal

band. Two methods currently exist for computing noise integrals of arbitrary order: the

residue theorem [7] and Lyapunov equations [37], both of which require extensive

arithmetical steps like matrix multiplication and inversion, as discussed in the next

section. For numerical integration, a circuit simulator can either use these methods or

numerical methods like the trapezoidal rule or second order backward difference

formula (gear2) [38]. However, if we want to perform symbolic integration to get

closed-form expressions of noise integrals, these two methods require exponentially

more time and memory as the order of the integral goes up. This renders these methods

impractical for higher orders of noise integrals.

Consequently, closed-form symbolic expressions of the noise integrals are

available only for the first few orders of noise transfer functions [39]. The absence of

closed-form expressions for higher orders makes it impossible to use noise in hand

analysis during the design of more complicated analog blocks. In addition, optimization

schemes for larger circuits like multiple-stage or multiple-feedback-loop amplifiers and

filters cannot be run without repetitive simulations involving lengthy numerical

methods for noise integration.

2.2 Previously Known Methods

As shown in (2-1), to find the total integrated noise, we need to compute

integrals of the form

. is an analytic function, a ratio of two

polynomials in , and .

. (2-2)

Because is an NTF in a realizable, stable circuit and because the integral of

is always bounded, and consequently the polynomials and

will have certain properties. A few properties that are relevant at this point are:

Chapter 2 - Symbolic Computation of Total Integrated Noise

11

1) The order of will be at-least 1 less than the order of . If not, then as

, the value of

will tend to a constant or tend to increase

indefinitely. In either case, the integral will not remain bounded.

2) will have no pole pairs on the imaginary axis because then, at the

corresponding frequencies, the value of will tend to infinity and the

integral will again not remain bounded.

3) cannot have a pole at DC because the integral again does not remain bounded.

In summary, we are interested in finding the integral of an NTF, as follows:

where

(2-3)

where is the non-zero coefficient of the highest order term in the denominator, .

The highest order term in the numerator is at most . Also, , which ensures

that there is no pole at DC.

Two previously known methods that can be used to compute the closed-form

symbolic expressions of are discussed next:

2.2.1 Residue Theorem

In [7], the residue theorem from complex analysis [40], is used to compute the

integral . The algorithm is depicted in (2-4) below for reference. More details can be

found in [7].

where

(2-4)

Chapter 2 - Symbolic Computation of Total Integrated Noise

12

The algorithm proceeds by first forming a companion matrix , and then passing

it as an argument to the two overloaded functions and as shown in (2-4).

represents the first derivative of with respect to its argument.

This method can efficiently provide the value of for numerical integration

(i.e., when the coefficients and are numbers, not symbols). For symbolic

integration, computing and would require us to find the power of

the matrix . This has an arithmetic complexity of . As we compute

higher powers of , the symbolic expressions become larger, and more memory is

required to store and manipulate the expressions. This means that as goes up, not

only are more arithmetic operations required, but the time taken for each operation also

goes up. As shown in Section 2.5 ahead, this method is impractical for symbolically

computing noise integrals with > 5.

2.2.2 Lyapunov Equations

The algorithm to compute the integral by solving a Lyapunov equation is

given below for reference. We first find a state space realization [41] of , i.e., find

the matrix , vector , vector , and the scalar , such that,

(2-5)

The elements of these matrices are functions of the coefficients and . After that, we

need to solve the Lyapunov equation for the matrix and then compute the

integral [37].

. (2-6)

. (2-7)

Note that because is strictly proper (i.e., the order of numerator is less than

the order of the denominator) [41]. There can be multiple realizations for the three

matrices, , , and for a given . Any realization can be used for finding the

noise integral through Lyapunov equations.

Chapter 2 - Symbolic Computation of Total Integrated Noise

13

For numerical integration (i.e., when the coefficients and are numbers, not

symbols), the Lyapunov equation in (2-6) can be efficiently and quickly solved using

the iterative eigenvalue algorithm, the QR decomposition [42], which reduces the

matrix to a complex Schur form [43]. However, for symbolic integration we would

have to solve the Lyapunov equation by first transforming it into a set of linear

equations in the unknown elements of the square matrix [44]. Solving

linear equations using Gauss elimination has an arithmetic complexity of , so

solving the Lyapunov equation (2-6) symbolically will have a complexity of . As

goes up, an increasing number of arithmetic operations are required, making this

method impractical for getting closed-form symbolic expressions for noise integrals for

higher values of , as shown in Section 2.5 ahead.

We thus see that it is generally difficult to obtain closed-form symbolic

expressions for higher order noise integrals in analog circuits using known methods. In

the next two sections, we develop a fast and direct method of computing these symbolic

expressions. We first develop a method for computing total integrated noise by

inspection in linear passive circuits and then extend it to find symbolic expressions for

noise integrals in an arbitrary circuit.

2.3 Computing Total Integrated Noise by Inspection in Linear,

Passive Networks

Consider a linear network made up of passive elements such as capacitors,

inductors, resistors, and transformers, with at-least one thermal noise generating

element. Let us say that we are interested in finding the PSD of the open-circuit noise

voltage or short-circuit noise current at a port of the network, due to all internal thermal

noise sources. Because the passive circuit is in thermal equilibrium, using the first and

second laws of thermodynamics, we can derive the Nyquist theorem which states that if

is the Laplace transform representation of the port impedance and

is the corresponding Fourier transform representation, then the two sided

PSD of the voltage noise at this port is given by [45]-[46],

Chapter 2 - Symbolic Computation of Total Integrated Noise

14

(2-8)

Similarly, if is the Laplace transform representation of the admittance at a port, the

two sided PSD of the current noise is given by

(2-9)

The mean squared value of the noise voltage or current can be computed by

integrating the respective PSD's,

(2-10)

In [47], it was shown that

(2-11)

While the above was derived for any or in [47], as shown in the extended

analysis in APPENDIX A, (2-11) works only when the functions and have

no poles at DC. The analysis in APPENDIX A also provides a complete proof of the

Bode theorem for impedance /admittance integrals, which states that for a general

or ,

(2-12)

In (2-12), the terms and have units of

capacitance, whereas the terms and have units of

Chapter 2 - Symbolic Computation of Total Integrated Noise

15

inductance. Let us call these capacitances, and respectively, and the

inductances, and respectively.

(2-13)

From (2-12) and (2-13),

(2-14)

For computing , the limit is taken as , i.e., as frequency tends

to infinity. As , for any inductor or resistor in the network, the product will

tend to infinity, whereas for a capacitor, the product will tend to its capacitance

value. We can thus open-circuit all the inductors and resistors, while keeping the

capacitors intact. Then, looking into the port, the net capacitance will be equal to .

Next, for computing , the limit is taken as , i.e., as frequency tends to

zero. As , for any inductor or resistor in the network, the product will tend

to zero, whereas for a capacitor, the product will tend to its capacitance value.

We can thus short-circuit all the inductors and resistors, while keeping the capacitors

intact. Then, looking into the port, the net capacitance will be equal to . Note that for

the cases where the port impedance has no poles at DC, the term , as

. In such cases, shorting all the inductors and resistors causes the port to be

shorted to ground. Consequently, for such cases, . Similarly, for computing

, we can short-circuit all the capacitors and resistors, while keeping the inductors

intact. Then, looking into the port, the net inductance will be equal to . For

computing , we can open-circuit all the capacitors and resistors, while keeping the

inductors intact. Then, looking into the port, the net inductance will be equal to , and

for the cases where the port admittance has no poles at DC, .

As an illustration of this method, consider the four circuits shown in Figure 2-1

(a)-(b) and Figure 2-2 (a)-(b). For the circuits in Figure 2-1 (a) and (b), we are

interested in finding , whereas for the circuits in Figure 2-2 (c) and (d), we are

interested in finding . Using the method described in the previous paragraphs, the

total integrated voltage and current noise is calculated by inspection for each of these

circuits.

Chapter 2 - Symbolic Computation of Total Integrated Noise

16

Figure 2-1. Computation of total integrated voltage noise by inspection.

(a)

R2

C2

C1C3C5R5

R4

L4

Nv

R2

C2

C1C3C5R5

R4

L4

(b)

R3

C2

C1

C∞n

R3

C2

C1

C0n

C∞n = C1 C0n = C1 + C2

Nv = kBTC∞n

1-

C0n

1 = kBT

C1

C2

C1 + C2

1

Z(s)

C∞n

C∞n = C1 + C2 + C3

C2C3

R2

C2

C1C3C5R5

R4

L4

C0n

C0n ∞

Nv = kBTC∞n

1-

C0n

1 = kBT

C2 + C3

C1C2 + C2C3 + C3C1

R3

C2

C1

Z(s)

Nv

Chapter 2 - Symbolic Computation of Total Integrated Noise

17

Figure 2-2. Computation of total integrated current noise by inspection.

(a)

Ni

(b)

L∞n

L∞n = L0n = L1

Ni = kBTL∞n

1-

L0n

1 = kBT

L3

1

Y(s)

L∞n

L∞n = L1

L0n

L0n ∞

Ni = kBTL∞n

1-

L0n

1 = kBT

1

L1

Y(s)

Ni

R2

C2

L3 L1

L1 R1

C2R2L2

R2

C2

L3 L1

L1 + L3

L1L3

L0n

R2

C2

L3 L1

L1 R1

C2R2L2

L1 R1

C2R2L2

Chapter 2 - Symbolic Computation of Total Integrated Noise

18

2.4 Track Mode Noise in Switched Capacitor Circuits

Switched-capacitor (SC) circuits are ubiquitous in today’s CMOS mixed-signal

ICs and are used primarily in filters, data converters, sensor-interfaces, and dc-dc

converters. Unfortunately, even though the basic theory of noise in SC circuits is

discussed in the literature, it is very intricate. The numerical calculation of noise in

switched circuits is very tedious, and requires highly sophisticated software [36]. In this

section, we use the visual inspection techniques developed in the previous section to

show how we can easily compute the total integrated track-mode noise in an SC

amplifier. As a vehicle that lets us explore our approach, we will utilize the commonly

used charge-redistribution track-and-hold amplifier (THA) stage [48], shown in Figure

2-3 below. and represent the sampling and feedback capacitances respectively,

while represents the parasitic capacitance at the input of the OTA.

Figure 2-3. Charge-redistribution track-and-hold amplifier stage.

Figure 2-4. Circuit model for noise analysis during track phase.

Cs

Cx

Cf

CL

φ1

φ1e

φ1

φ2

φ2

φ2

Vin

Vout

φ1

φ1e

φ2

VXVs

Vf

Cs

Cx

Cf

Vx

Rxn

Rsn

Rfn

Vxn2

Δf= 4kBTRxn

Vfn2

Δf= 4kBTRfn

Vsn2

Δf= 4kBTRsn

Vs

Vf

qx

Chapter 2 - Symbolic Computation of Total Integrated Noise

19

A circuit model for noise analysis during tracking is shown in Figure 2-4, which

contains only the relevant circuit elements and noise sources that are involved in this

phase. The resistors , , and , represent the ON resistances of the and

switches connected at nodes , and .

At the end of the track mode, the total noise charge at node is frozen and

then this charge is redistributed onto the feedback capacitor in the clock phase. The

output noise contribution can then be written as

(2-15)

In summary, to find the contribution from the track phase referred to the output of the

THA, we are primarily interested in finding the mean square value of this noise charge

.

The total charge at node can be written as follows:

(2-16)

The mean square value of the charge can thus be written as

(2-17)

Note that while the noise source voltages , and are statistically independent

and thus uncorrelated, the noise voltages at the nodes , and are correlated.

Consequently, the cross-correlation terms ,

and are non-zero. These cross-

correlation terms can be found by first finding the mean square noise voltage difference

between the three nodes, , and as follows:

(2-18)

Chapter 2 - Symbolic Computation of Total Integrated Noise

20

Figure 2-5. Computing mean squared noise voltages by inspection.

Cs

Cx

Vx

RxnRsn

Vs

Cf

Rfn

Vf

Cs

Cx

Vx

RxnRsn

Vs

Cf

Rfn

Vf

Cs + Cx

Cs Cx

Cf + Cx

Cf Cx

Finding C∞n

Cx

Finding C0n

∞ ∞ ∞

Vs2 = kBT

Cs + Cx

Cs Cx

Vx2 = kBT

1

Cx

Vf2 = kBT

Cf + Cx

Cf Cx

Cs

Cx

RxnRsn

Vs

Cf

Rfn

Vf

Finding C∞n Finding C0n

(Vs – Vf)

2 = kBT

Cs + Cf

Cs Cf

Cs + Cf

Cs Cf

Cs

Cx

RxnRsn

Vs

Cf

Rfn

Vf

Cs

Cx

RxnRsn

Vs

Cf

Rfn

Vx

Finding C∞n Finding C0n

(Vs – Vx)

2 = kBT

1

Cs

Cs

Cx

RxnRsn

Cf

Rfn

∞Cs

Vs Vx

Cs

Cx

RxnRsn

Vf

Cf

Rfn

Vx

Finding C∞n Finding C0n

(Vf – Vx)

2 = kBT

1

Cf

Cs

Cx

RxnRsn

Cf

Rfn

∞Cf

VfVx

Chapter 2 - Symbolic Computation of Total Integrated Noise

21

From (2-17) and (2-18),

(2-19)

The mean squared voltages in (2-19) can be found by using the method

described in the previous section, as shown in Figure 2-5. Finally, from (2-19) and

Figure 2-5, we get

(2-20)

which matches the result obtained by a complete symbolic integration of the power

spectral densities at the different voltage nodes performed in [48]. Using the method

described in Section 2.3, the same result can be derived by visual inspection and a small

number of algebraic steps.

2.5 Exact Closed-Form Expressions for Total Integrated

Noise in Arbitrary Circuits

While the methods developed in the previous two sections can provide quick

noise performance estimates by visual inspection, they work for passive circuits.

However, almost all practical circuits of interest contain at least one transistor and the

corresponding linearized circuits contain active elements such as controlled current

and/or voltage sources. Such active elements provide energy and hence an active circuit

is not in thermal equilibrium. Thus, unlike the case of passive circuits, Nyquist's

theorem for the PSD of the output noise cannot be applied to the general case of active

circuits.

In this section, we show how we can indirectly use the Nyquist theorem in any

general circuit to compute closed-form expressions for noise integrals. In order to do so,

we make the following important observation: As shown in APPENDIX A, (2-11) was

derived for an impedance expression using the initial value theorem and the

relationship between inverse Laplace and inverse Fourier transforms. We saw that this

was possible only when the function has no poles at DC, has no poles on the

Chapter 2 - Symbolic Computation of Total Integrated Noise

22

imaginary axis, has only complex conjugate pole pairs with negative real parts or poles

on the negative real axis, has zeros anywhere, and the number of zeros is at least one

less than the number of poles. This means that for any analytic function that has

the above properties, we can write

(2-21)

where we have used the fact that is an even function of in going

from (2-11) to (2-21).

In Section 2.2, we saw that in an arbitrary, active or passive, linear, analog

circuit, we are faced with finding the closed-form expression for an integral of the

kind given in (2-3). If we now assume that for such an NTF (with , , ,

and conditions on poles and zeros as mentioned in Section 2.2), it is always possible to

find a rational function (with the properties that enable (2-21) to hold), such that

the following equality holds:

(2-22)

then we can use (2-21) and (2-22) to get closed-form expression for the integral . We

next show how we can find the function . The left hand side of (2-22) can be

written as,

(2-23)

Similarly, the right hand side of (2-22) can be written as,

(2-24)

From (2-22), (2-23), and (2-24),

(2-25)

(2-25) will always hold if the following equation holds:

Chapter 2 - Symbolic Computation of Total Integrated Noise

23

(2-26)

Next, assume that and are the numerator and denominator

polynomials of , respectively.

(2-27)

From (2-26) and (2-27):

(2-28)

In (2-28), we know and , but we do not know and . Let us say we

take equal to . The poles of then become the same as the poles of .

The limitations we imposed on the poles of will then also apply on the poles of

. However, these are the same properties (no pole at DC, no pole pairs on the

imaginary axis, only complex conjugate pole pairs with negative real parts or poles on

the negative real axis), that enable (2-21) to hold. This makes a good

choice. Thus,

(2-29)

From (2-28) and (2-29),

(2-30)

(2-30) can be solved to find the unknown polynomial by using the known

polynomials and . The order of is and the number of zeros in

has to be at-least one less than the number of poles. The order of will thus be

. We can thus write,

(2-31)

where the m coefficients are unknown.

Finally, from (2-3), (2-21), (2-22), (2-30), and (2-31),

Chapter 2 - Symbolic Computation of Total Integrated Noise

24

(2-32)

(2-32) shows that for finding the integral we need to find only the coefficient

and not the entire . As shown in APPENDIX B, where and

are two matrices constructed from the coefficients and as shown in

(2-33) and (2-34) below.

(2-33)

(2-34)

where denotes the largest even integer smaller than or equal to and

for and Also, denotes the modulo operation, i.e.,

for even, and for odd.

Table 2-1 provides expressions for the matrices and , and provides the

closed-form expressions for 1, 2, 3, 4, 5.

Chapter 2 - Symbolic Computation of Total Integrated Noise

25

Table 2-1. Closed-form symbolic expressions for .

Chapter 2 - Symbolic Computation of Total Integrated Noise

26

We see that our method involves computing the determinant of matrices,

the elements of which are the coefficients of and . The arithmetic complexity

of our new method is thus . This makes our method faster than the Lyapunov

equation method [37], which has an arithmetic complexity of as discussed in

Section 2.2.2. Additionally, our method is direct because there is no need to store long

symbolic expressions intermediately.

In order to test the relative and absolute speeds of the three aforementioned

methods, symbolic integrations were performed with the aid of the MATLAB Symbolic

Math Toolbox [6], on a CPU with a quad-core processor running at 2.2 GHz with 4 GB

available memory. For small values of , when the integration completes in a short

amount of time (less than a few seconds), multiple runs were done and the time taken

was averaged. Figure 2-6 shows plots of time taken versus the order for the three

methods.

As shown in Figure 2-6, the residue theorem method is the slowest and fails to

execute on our machine for 5, because of in-sufficient memory. This is because the

intermediate symbolic expressions become extremely large, even for relatively small

values of . The performance of the Lyapunov equation method is better, but it takes

greater than 1 day to perform symbolic integrations for . On the other hand, the

proposed method is the fastest and takes no more than a minute to perform symbolic

integration of noise integrals up to an order of = 13, taking much less time than the

Lyapunov equation method for higher values of .

Chapter 2 - Symbolic Computation of Total Integrated Noise

27

Figure 2-6. Comparison of time taken by three methods of symbolic integration.

2.6 Conclusion

In this chapter, we showed how we can visually compute the total integrated

noise in passive circuits and used the method to compute the track-mode noise in a

switched-capacitor stage. The method was then extended to show how we can compute

the exact closed-form expressions for noise integrals of any arbitrary order in an active

or passive analog circuit. The end expressions derived in this work, shown in Table 2-1,

had been previously derived in a similar form but using a different mathematical

approach, in a study of noise in nonlinear servomechanisms [49], and were pointed out

by the reviewers of the IEEE Circuits and Systems Society. The results presented here

have been derived independent of that work.

Closed-form expressions developed here for total integrated noise can now be

used for hand analysis during design. In addition, hitherto under-utilized optimization

algorithms can now be used for complete and exact noise optimization of analog

circuits like multiple stage or multiple feedback loop amplifiers and filters as shown in

the next two chapters.

1 3 5 7 9 11 130.01

1

100

10,000

1,000,000Residue theorem Lyapunov equation This work

1 3 5 7 9 11 130.01

1

100

10,000

1,000,000

Tim

e t

ake

n [

seco

nd

]

1 minute

1 hour

1 day

m (Order of Noise Integral)

Chapter 2 - Symbolic Computation of Total Integrated Noise

28

29

CHAPTER 3

Settling Time and Noise Optimization of a

Three-Stage OTA

3.1 Motivation

Fast, high-gain operational-transconductance-amplifiers (OTAs) are an integral

part of switched-capacitor (SC) circuits. Figure 3-1 shows a model of a typical SC gain

stage in the charge redistribution phase.

Figure 3-1. Model of an SC gain stage during charge redistribution.

The OTA has been traditionally implemented using a cascade of two stages to

provide a high gain. However, as shown in Figure 3-2, the intrinsic gain of the

transistors used in these amplifiers has reduced dramatically due to technology scaling

( 15 for 90-nm CMOS). In addition, due to voltage headroom constraints, it has

become increasingly difficult to use cascoding as a solution to this problem. Thus, as an

alternative, it is attractive to consider cascades of more than two common-source stages

to achieve high DC gain [50]. A popular topology for this is the nested-Miller-

compensated three-stage OTA with a nulling resistor [51], [52].

CF

CF

CS

CSCL

CL

VidVod

Vop

VomVinp

Vinm

Vim

Vip

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

30

Figure 3-2. Transistor intrinsic gain across technology nodes.

While the design of such a three-stage OTA is relatively straightforward for

continuous-time applications, the presence of multiple poles and zeros makes the design

challenging when accurate and fast transient settling is required. Improving the settling

performance by phase-margin adjustments has been proposed in the literature. However,

for three-stage amplifiers, better phase margin does not always imply faster settling [53].

Analytical and numerical methods have been developed that find the best pole-zero

locations for optimal settling based on open-loop or closed-loop analysis [53], [54].

However, such analyses do not include all effects, such as the impact of capacitive

feedforward, on the settling performance.

In addition, all previously published three-stage amplifier design methods

address only the settling performance. Transistor thermal noise from the three stages

and noise from the nulling resistors undergo frequency-dependent peaking based on the

pole-zero placements in their respective noise transfer functions (NTF's). Depending on

how much transconductance is allocated to each stage and the values of the

compensating capacitors, there can be multiple sets of closed-loop transfer functions

and internal NTF's. Each design in this set might have different total integrated noise,

while meeting the specifications for settling time and power consumption. This means

that a design with optimal settling time might not have the best noise performance.

In this chapter, we use closed-form symbolic expressions for total integrated

noise, settling time, and power consumption to form an optimization program [55], [56].

We can optimize any one of the performance metrics while putting a constraint on the

other metrics. As an example here, the optimization program is set up to minimize the

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

31

total integrated noise while putting a constraint on the power consumption and settling

time. In the next sections, we describe the three-stage OTA specifications, schematic, a

behavioral model, and present the design equations and constraint formation steps. We

then present the final design, and compare the calculated and simulated results in

Section 3.3.

3.2 Design and Optimization Framework

The transistor level implementation of the three-stage OTA, along with the two

common-mode feedback stages, is shown in Figure 3-3. In order to illustrate the design

and optimization of the OTA, we start with a list of specifications summarized in Table

3-1. These assumed values are for a typical SC gain stage clocked at 200 MHz, which

finds applications in pipelined ADCs and SC delta-sigma modulators. The dynamic

error specification of 0.1% implies that the output should settle to within +/- 0.1% of the

final steady state value within 2.5 ns (1/2 clock cycle). For simplicity, it is assumed that

the OTA does not slew in the initial part of the transient. Slewing is well understood

and its effect can be easily incorporated in the optimization if desired. Thus, we focus

here on optimizing only the linear settling part of the transient, which is analytically

more complex. Finally, note that the ideal closed loop gain of the switched-capacitor

amplifier is approximately given by and thus for the desired closed-

loop gain.

Technology 90-nm CMOS Dynamic Error, < 0.1 %

1 V Settling Time, < 2.5 ns

Power, < 5 mW Closed-Loop Gain 2

Load Cap., > 200 fF Total Integrated Noise Minimize

Table 3-1. Design specifications for SC gain stage.

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

32

Figure 3-3. Transistor-level implementation of the OTA (bias network not shown). All NMOS

bodies are tied to ground, all PMOS bodies are tied to .

3.2.1 Model Description

The transistors shown in Figure 3-3 are sized using the methodology [9].

Based on power efficiency, high , good linearity and high requirements, the

values and channel lengths of the transistors shown in Figure 3-3 are fixed a

priori to the values shown in Table 3-2. The load transistors and do not

contribute to transconductance, but contribute to the noise of the amplifier. In order to

reduce the noise contribution from the load transistors, their corresponding are

scaled down by selecting a smaller of S/A, as shown in Table 3-2. Note that

Vb1

Vb2

Vb4

Vb3 Vb5

Cm2Cm2

RmRm

Cm1

Vinp Vinm

V1m

V1p

V2p

V2m

VomCm1 Vop

Stage 1 Stage 2 Stage 3

M1n

M1p

M1tail

M2tail

M2n

M2p

M3n

M3p

M1cmfb

M2cmfb

V1cmfb

V2cmfb

Vb6

Vb7

Vc1

RaRa

Vb8

RaRa

V1m V1p

Vom Vop

Vb9

Vc2

CMFB 1 CMFB 2

V1cmfb

V2cmfb

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

33

these parameters are subject to optimization themselves, but are being viewed as

constants here for simplicity.

Device , Length

, , 18 S/A, 140 nm

, , 7 S/A, 240 nm

, 10 S/A, 240 nm

, 10 S/A, 240 nm

Table 3-2. design choices.

A lookup table based methodology, as explained in [57], can be used in

conjunction with the optimization routine used in this work in order to further improve

the final design. In addition, in order to take into consideration the effect of process and

temperature variations, we can either tighten the specifications of the design, or

generate the characterization plots using SPICE models at the worst process

corner [57].

The single-ended behavioral model of the SC amplifier, together with the OTA

and the source and feedback capacitors, is shown in Figure 3-4. Also shown are the

definitions of to , which model the sum of the parasitics in the circuit.

represent the DC gain of the three stages and are assumed to be equal to 8 based on the

of the transistors, which themselves are fixed because the are fixed. Note

that the exact value of these DC gains do not matter much in the closed loop analysis so

long as the total DC gain of the three-stage amplifier is reasonably large (say ).

All the parasitics scale with the transistor widths, which themselves scale with

the respective transconductances . Therefore, fixing the s enables us to

express all bias currents, model parameters, constraint functions, and the objective

function, in terms of just the following seven circuit variables: , , , ,

, , and .

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

34

Figure 3-4. Behavioral model of nested-Miller-compensated OTA in an SC amplifier.

3.2.2 Equations and Constraints Formation

A. Inequality Constraints - Power Consumption and Load Capacitance

The power consumption of the circuit is given by

. (3-1)

The factor (taken to be 0.2), is used to take into consideration the extra power

consumed in the bias and common-mode feedback (CMFB) branches. We constrain the

power consumed to a certain value , forming the first inequality constraint,

. (3-2)

The second inequality constraint is related to the load capacitance. We want the

OTA to drive a certain target load capacitance . We can always add more

capacitance to improve noise, but the minimum load capacitance is . Thus,

(3-3)

C1 C2

Cin

C3R1

Vin

R3R2

V1 V2 Vo

gm1Vin gm2V1 gm3V2

Cm1

Cm2 Rm

* * *

*

Vx

in12 in2

2in3

2

vnm2

CS

CF

Vi

R1 = A1/gm1 R2 = A2/gm2 R3 = A3/gm3

Rm = 1/gm3

Cin = Cgg(M1n)

+ A1Cgd(M1n) C1 = Cgg(M2p) +

A2Cgd(M2p) +

Cdb(M1n) + Cdb(M1p)

C2 = Cgg(M3n) +

A3Cgd(M3n) +

Cdb(M2p) + Cdb(M2n)

C3 = CL +

Cdb(M3n) + Cdb(M3p)

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

35

B. Inequality Constraints - Settling Time

The closed-loop transfer function of the three-stage OTA, , has a dominant

pole, a complex conjugate pole pair, a high frequency pole, a right half plane (RHP)

zero and three left half plane (LHP) zeros. Normalizing the poles and zeros to the

dominant pole, we can write in terms of the following eight pole-zero variables,

, , , , , , , and , as shown below:

(3-4)

The variable is negative, while all other variables are positive. The normalized step

response of the system, illustrated in Figure 3-5, is assuming

no slewing at the beginning.

Figure 3-5. Normalized step-response.

The dynamic error is given by . Using partial fraction

expansion, the expressions for and are derived as shown in (3-5) and

(3-6),

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

36

(3-5)

(3-6)

where it is assumed that 0 . The coefficients and thus , are functions

of the eight pole-zero variables.

The design specification states that the dynamic error should be within 0.1 %

for 2.5 ns. This can be written as a semi-infinite constraint [10]:

-0.001 0.001 for 2.5 ns. (3-7)

Without any additional constraints on the pole-zero positions, the optimization

loop might return a design with the zeros very close to the poles , , and

. While such pole-zero cancellations can improve , slight deviations in the

pole-zero positions because of process variation can lead to a sharp increase in [58],

[59]. In order to have a robust system, we should have the poles and zeros spaced apart

by a certain factor. The analysis in [59] shows that deviations in the pole-zero positions

do not severely impact , when the poles and zeros are spaced apart by a factor of

more than 1.5. In our case, we constrain this spacing factor to a practical value of 2.

This is incorporated in the design by adding the following constraints to our

optimization loop,

for = 1..4. (3-8)

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

37

C. Equality Constraints - Mapping Pole-Zero Positions to Circuit Design Variables

In (3-4), is expressed in terms of the eight pole-zero variables: , , ,

, , , , and . By performing symbolic analysis on the behavioral model

of Figure 3-4, we can also express in terms of the seven circuit variables: ,

, , , , , and . Equating the corresponding coefficients of the

numerator and denominator polynomials of these two expressions for , we get eight

equations in terms of fifteen design variables,

for (3-9)

D. Optimization Objective — Total Integrated Noise

The thermal noise PSD's of the four noise sources in Figure 3-4 are,

. (3-10)

where are the excess noise factors of the respective transistors. Because we have

fixed the lengths of the transistors, the value of the noise factors can be extracted from

the technology characterization plots. We find that for the 140-nm NMOS devices,

1.25, and for the 240-nm NMOS device, 1.19. Similarly, for the

140-nm PMOS devices, 1.97, and for the 240-nm PMOS devices,

1.33. Let us label the NTF's of these noise sources to the output as , ,

and . The total integrated noise of the complete amplifier can then be

written as

(3-11)

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

38

The total noise is doubled because the circuit is symmetric and the output is sensed

differentially [7]. The four NTF's can be computed by symbolic analysis of the circuit

in Figure 3-4. The coefficients of the numerator and denominator polynomials of these

NTF's will then be functions of the circuit design variables.

According to linear circuit theory [60], the circuit poles are the inherent natural

resonant frequencies of the circuit, and hence all transfer functions in the circuit will

have the same set of poles. This means that the poles, and hence the denominator

polynomials of the NTF's will be the same as that of . This helps us reduce the size

of the symbolic expressions in the four NTF's because the denominator polynomials in

the NTF's can be replaced by the simpler denominator polynomial of given in

(3-4). Thus, we can write the coefficients of the numerator polynomials of the NTF's as

functions of the seven circuit design variables and the coefficients of the denominator

polynomial as functions of the eight pole-zero variables.

The next step is to symbolically integrate the squares of the magnitudes of these

NTF’s across all frequencies. Using the formula in Table 2-1 with = 5 (because all

NTF's are 5th

order), the complete closed-form symbolic expression for total integrated

noise in (11) can be expressed as a function of the fifteen design variables,

( ). (3-12)

3.2.3 Running the Nonlinear Constrained Optimization Program

We use the Optimization Toolbox in MATLAB [10] to formulate our noise

optimization loop as a constrained non-linear optimization program as shown in (3-13)

below.

min.: ( )

subj. to:

for

-0.001 0.001 for ns

for = 1..4 (3-13)

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

39

Closed-form symbolic expressions of the constraints and the objective, and the

gradients of the objective are passed on to the fseminf function and the interior-point

algorithm is used for optimization [10]. As shown in Figure 3-6, design heuristics are

used to provide a good initial starting point. For example, we can assume that the first

stage will contribute more to the total integrated noise of the amplifier than the

consecutive stages. Consequently, in the initial design point, more transconductance is

allocated to the first stage than the other stages.

Figure 3-6. Running the optimization program.

Running the optimization loop takes a short amount of time. Note that the

constraint and objective functions are nonlinear, non-convex [10] functions of the

design variables. These functions can have multiple local minima values, and

consequently, the optimization loop might converge to a local minima, instead of the

global minima. One way to check if we have a solution that is the globally optimal

solution or is close to the globally optimal solution, is to provide a different starting

point and then check if the loop converges to the same optimal point. Doing this is fast,

because each run of the loop does not take a lot of time, unlike the case where a

simulator is needed within the optimization loop. Additionally, we can also confirm the

optimality of the final design point by checking if it meets the following necessary

condition: say that the optimization program shown in (3-13) leads to a design point

that exhibits a total integrated noise . We then construct another optimization

program just like the one in (3-13), but now minimize the total power consumed while

constraining the total integrated noise to . If this new optimization program also

leads to the same design point and has a total power consumption of , then such a

final design point can be assumed to be the globally optimal solution or close to it.

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

40

3.3 Final Design

The optimized design variables and parameters are shown in Table 3-3. The

complete transistor level circuit is simulated using Spectre at the transistor level and the

achieved performance is given in Table 3-4. We also compare the Spectre simulation

results to our behavioral model calculations in order to confirm the accuracy of the

model. As shown in Figure 3-7, the simulated transient response matches the calculated

transient response well.

In Figure 3-8, the calculated and simulated noise PSD's from the four noise

sources are plotted, which again shows close matching. Finally, total integrated noise

contributions from the four noise sources, calculated and simulated, are tabulated in

Table 3-5, showing good matching between the behavioral model and the transistor

level circuit. The difference between the assumed DC gain from the second stage and

the actual DC gain causes the discrepancy between the calculated and the simulated

low-frequency second-stage noise PSD. However, the noise PSD matches well at high

frequency, and consequently, the calculated and the simulated total integrated noise

from the second stage is nearly the same.

98 μm / 0.14 μm 1 V 0.507 pF

53 μm / 0.24 μm 900 μA 1.86 pF

44 μm / 0.24 μm 262 μA 385 MHz

44 μm / 0.24 μm 938 μA 0.4577

72 μm / 0.14 μm 16.03 mS 0.5257

7 μm / 0.24 μm 4.86 mS 0.0677

30 μm / 0.24 μm 16.94 mS -0.2629

30 μm / 0.24 μm 59 Ω 0.1682

103 μm / 0.14 μm 1.89 pF 0.1682

56 μm / 0.24 μm 1.26 pF 0.1682

Table 3-3. Design parameters.

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

41

Specification Performance

Power Consumption < 5 mW 5.2 mW

Settling Time < 2.5 ns 2.53 ns

Total Integrated Noise Minimize 240 μVrms

Table 3-4. Performance summary.

Figure 3-7. Settling to within dynamic error.

Figure 3-8. Comparison between calculated and simulated noise PSDs.

Chapter 3 - Settling Time and Noise Optimization of a Three-Stage OTA

42

Noise Source Total Integrated Noise

Calculated Simulated

First Stage Transistors 166 μVrms

163 μVrms

Second Stage Transistors 132 μVrms 137 μVrms

Third Stage Transistors 93 μVrms 91 μVrms

Nulling Resistors 70 μVrms 66 μVrms

All sources 241 μVrms 240 μVrms

Table 3-5. Comparison between calculated and simulated total integrated thermal voltage noise.

3.4 Conclusion

This chapter described the design and optimization of a three-stage amplifier for

SC applications. Previously known design methods for three-stage amplifiers usually

optimize just one of the design metrics while neglecting other important metrics. In our

approach, one of the design metrics, the total integrated noise, is optimized given a

constraint on the other design metrics. The procedure can be modified to optimize other

specifications, say the settling time, given a constraint on other metrics, or for multi-

objective optimization. The design and optimization approach can be applied to other

topologies as well. Since the constructed behavioral model includes all relevant device

parasitics, there is no need to interface with a circuit simulator. As we have shown, this

leads to a fast optimization program, while maintaining close matching to circuit-level

simulation.

43

CHAPTER 4

Design and Optimization of Continuous-

Time, gm-C and Active-RC Filters

4.1 Motivation

Filters form an integral part of the analog front-end. Continuous-time (CT)

analog filters are important components of high-speed wireless and wireline systems in

which they are used to select a particular channel of interest out of a densely packed

spectrum. Filters are also used to block interferers, and to provide anti-alias filtering

before the subsequent analog-to-digital conversion stages [30]. Two common

topologies used to implement CT filters are the gm-C topology, which uses

transconductance cells and capacitors, and the active-RC topology, which uses

operational-amplifiers (op-amps), resistors and capacitors [61].

The challenges in the design and integration of high-performance CT analog

filters in modern day electronic systems are well documented and can be briefly listed

as follows [61]:

Noise: Resistors and active devices such as transistors generate thermal noise. The

root mean squared noise voltage, found by integrating the noise generated by the

components in the filter across all frequencies, effectively determines the dynamic

range of the analog filter. Any signal smaller in strength than this noise voltage floor

cannot be intelligibly distinguished from the noise, rendering the filter ineffective

for smaller amplitudes. This problem is further aggravated at smaller technology

nodes where sub-1V power supplies are used, limiting the available signal swing. In

such a case, the noise floor of the filter needs to be further reduced to achieve large

dynamic range.

Linearity: Analog filters are typically implemented as cascades or nested loops of

multiple stages, where each stage’s output driver is implemented using transistor

stages. As the voltage swings in these transistor stages increase, nonlinearities are

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

44

introduced in the signal chain. Consequently, the voltage swings are limited to a

fraction of the supply voltage rails. This effectively reduces the dynamic range of

the filter.

Area: Passive components, like resistors and capacitors, in active-RC filters and

capacitors in gm-C filters, occupy precious chip-area. The reduction of area for

analog filter designs usually implies a tradeoff, most common of which is an

increase in noise.

Power consumption: The amplifier stages used in active-RC and gm-C topologies

consume power. In general, the filter’s power consumption scales with its

bandwidth and dynamic range.

Frequency response stability: Any variations in the frequency response

characteristics like cut-off frequency, pass-band ripple, stop-band attenuation, etc.

can cause system performance deterioration. Such variations are induced by

temperature change and fabrication tolerances. In robust designs, these variations

are typically controlled to be less than 1%. This is usually done by choosing

topologies with nested feedback loops, like gm-C and active-RC topologies, which

are less sensitive to parameter variation than other topologies that use cascaded

stages [30]. In this chapter, we will focus only on gm-C and active-RC topologies.

High-frequency performance: Analog filters are impacted by parasitic capacitances

as well as finite transistor fT, which can cause the realized filter’s frequency

response to deviate from the desired ideal response.

In summary, CT analog filter design problems seek to find robust, low-area,

low-power circuit solutions that maximize dynamic range. There are three steps

involved in finding the right solution as illustrated in the decision diagram shown in

Figure 4-1.

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

45

Figure 4-1. Analog filter synthesis steps.

As a first step, starting from a set of filter specifications, one out of different

kinds of transfer functions like Butterworth, Chebyshev or elliptic functions is chosen

[30]. Each function differs in the way the corresponding poles and zeros are placed.

Consequently, each function has different roll-off characteristics, pass-band ripple,

group-delay response, and the choice of a particular kind of function depends on system

specifications. This choice can be done by trial-and-error runs on MATLAB [62].

Typically, in applications where the phase response is secondary, an elliptic transfer

function is chosen because for a given stop-band attenuation and transition-band

specification, an elliptic transfer function has the smallest order, and consequently the

smallest number of active components [63]. Once a suitable transfer function is chosen,

the corresponding LC ladder filter realization is implemented using filter synthesis

tables [64] or MATLAB [62]. Finally, the node voltages and branch currents in the LC

ladder filter circuit are scaled and mapped to the corresponding voltages and currents in

a gm-C or active-RC circuit, determining a set of component values for the resistors,

capacitors and transconductors.

In this chapter, we deal with the last part of the process mentioned above,

namely, developing a design and optimization framework that can find the optimal

component values in a CT analog filter from a given LC ladder realization. Before we

Filter Specifications

Transfer Function Derivation

Derived from system specifications / standards.

Choosing one of Butterworth, Chebyshev I & II, elliptic

functions so that required specifications are closely met.

LC Ladder Filter Realization

A passive, LC-based equivalent filter is constructed based on

pole-zero placements in chosen transfer function.

Heuristics, MATLAB

Standard tables, MATLAB

Transistor Level Design

The node voltages and branch currents in the LC filter are scaled

and mapped to corresponding voltages and currents in active-

RC or gm-C structure.

Signal-flow-graph manipulation, optimization

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

46

describe the proposed approach, a brief overview of prior work done in analog filter

design and optimization is provided in the next sub-section.

4.1.1 Prior Work in Filter Design and Optimization

The prior literature on CT filter design and optimization can be broadly

classified into two kinds of approaches: heuristics-based and mathematics-based.

Heuristics-based: In this approach, simplifying assumptions are made about the

components that reduce the scale of the underlying node-equations. As an example,

in an active-RC topology, either all resistors are kept equal to each other as shown

in [65], or all capacitors are kept equal to each other as shown in [66]. Similarly in a

gm-C topology, either all transconductors are kept the same as shown in [67] and

[68], or all capacitors are kept the same as shown in [69]. Thereafter, node-voltage

scaling [70] is performed, where each stage’s output signal-levels are maximized, in

an effort to maximize the signal handling capacity of the filter. The noise

performance of the filter is typically not dealt with in such approaches, rendering the

final design sub-optimal.

Mathematics-based: In mathematics-based approaches, attempts are made to

formally derive various equations for the different performance metrics and make

them a part of the design and optimization routine. In [32] and [33], theoretical

bounds for noise in active-RC and gm-C topologies have been derived. In [71], a

method for optimally reducing die area in a given active-RC filter design, while

maintaining the filter’s designed signal-to-noise ratio (SNR), frequency response,

and topology is discussed. However, it makes three simplifying assumptions that are

not always applicable in the general case. Firstly, the noise from op-amps is not

considered. Secondly, noise transfer functions from each resistor source to the

output are assumed to be like an ideal brick-wall response at the filter’s cut-off

frequency. This neglects any peaking that is introduced by the poles and zeros in the

individual noise transfer functions. Lastly, nonlinear constraints that relate the

resistor and capacitor values to the poles of the filter transfer function are linearized

using a first order Taylor series expansion. This is done so that a linear-

programming based optimization routine can be used. In [72], [73], [74] and [34],

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

47

matrix methods are developed to find an optimal set of component values that

maximizes the dynamic range for a given power consumption budget. While these

methods truly give an optimal design, they do not work for the cases where there are

finite zeros in the filter transfer function. An example of such a transfer function is

the elliptic function, which is also the most popular and economical filter transfer

function used most often for implementing active filters. Furthermore, op-amp noise

in active-RC filters is again neglected. Finally, these methods cannot be used with

any additional constraints that the designer would want to include. These include

constraints on area, and constraints that enforce the use of integral multiples of unit

resistors or unit capacitors or unit transconductors for better immunity to component

variation.

In this chapter, we propose a design and optimization methodology that

overcomes many of the shortcomings described above. In particular, we describe the

optimization of the dynamic range, power consumed and area, in gm-C and active-RC

topologies that implement a 5th

order elliptic transfer function. Closed-form symbolic

expressions are used for noise, power and area and the optimization problem is

formulated as geometric programs (GP) and mixed-integer geometric programs (MIGP).

As a first step, we describe GPs and MIGPs in the next section.

4.2 Geometric Program (GP) and Mixed-Integer Geometric

Program (MIGP)

Let be a real-valued function of real, positive variables . This

function is called a posynomial function if it has the form given below:

(4-1)

where and . When is called a monomial function. A geometric

program (GP) [75] is an optimization problem of the form

minimize:

subject to:

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

48

(4-2)

where and are posynomial functions and are monomial functions. A GP is a

special type of optimization problem where the objective function and the constraints

are convex functions [30]. The most important feature of GPs is that they can be

globally solved with great efficiency. The optimization algorithm also determines

whether the problem is infeasible (i.e., no design can meet all constraints). Given the

global nature of the solution, the starting point for the optimization algorithm does not

have any effect.

A variant of the GP described above is a mixed-integer GP (MIGP) [75]. In an

MIGP, we have a GP with the additional constraint that some of the variables lie in

some discrete set, such as a set of integers. An MIGP can thus be represented as:

minimize:

subject to:

, (4-3)

where denotes the set of natural numbers, i.e., positive integers.

MIGPs are in general hard to solve, and all methods for solving them make

some compromise compared to the methods for GP. Heuristic methods attempt to find

good approximate solutions quickly, but cannot guarantee that the solution found is

optimal. Global methods always find the global solution, but can take a long time to run.

In this work, we use a common global method that is based on branch-and-bound

algorithms [76]. This method is non-heuristic, in the sense that it maintains a provable

upper and lower bound on the globally optimal objective value, and ultimately finds the

true global solution.

A typical branch and bound method works as follows [75]: We remove the

integer constraints and let the solver solve this relaxed GP. This gets us to a lower

bound on the optimal value of the MIGP, and we use some heuristic (such as rounding)

to find a locally optimal approximate solution. If it is feasible, and its objective value is

close enough to the lower bound, we can take it as a guaranteed nearly optimal solution,

and quit. If this is not the case, we choose an integer variable and branch. To do this

we choose some integer , and observe that if is the optimal value of (for the

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

49

MIGP) then either or

. We form two children MIGPs, from the

original one, by adding these two constraints to the original problem. We then repeat the

above for each of these children. Eventually, the constraints on the integer variables

become equality constraints, at which point we have GPs and the relaxation yields the

exact global solution. The hope is that this happens within a reasonable number of steps.

Branch and bound algorithms differ in the heuristics used to choose a variable to branch

on, and on which children MIGPs to process at each step.

As shown ahead, given our choice of design variables and equations for the

design of the gm-C and active-RC filter, the functions for total integrated noise, area and

power consumption turn out to be posynomial functions of the design variables. This

enables the formulation of the optimization problem as a GP. Instead of implementing

our own algorithm for solving the GP, we use yalmip [77], [78], a free MATLAB

software package for specifying and solving convex programs.

4.3 Design and Optimization Framework

4.3.1 Elliptic Filter Structure

The filter specifications used for demonstrating the proposed design and

optimization methodology are listed in Table 4-1. Note that gm-C topologies are better

suited for filters with higher cut-off frequencies, and consequently, different cut-off

frequency specifications are used for the two topologies.

Transfer Function 5th

Order Elliptic Pass-Band Ripple < 0.5 dB

Cut-Off Frequency, 30 MHz (gm-C)

1 MHz (active-RC) Pass-Band Gain 0 dB

Stop-Band Edge

Frequency 1.33

Source / Load

Resistance 500 Ω

Table 4-1. Filter design specifications.

Figure 4-2 shows the passive LC equivalent filter with component values

( - ) annotated. These are found by first finding the normalized

component values using MATLAB [62], and then scaling them for frequency and

source/load resistances. Note that the input is gained up by a factor of 2 because the

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

50

passive structure by itself has a DC gain of 1/2. The internal node voltages ( )

and currents ( ) are also annotated. Note that denotes the output

voltage. Figure 4-3 shows the corresponding single ended gm-C implementation of the

filter. We see that it uses 11 transconductor stages and 7

capacitors . Figure 4-4 shows the single-ended active-RC implementation.

The corresponding differential implementations are shown in Figure 4-5 (a) and (b).

The internal node voltages in both topologies are annotated (

) while

the output node voltage is annotated as . Note that both the representations shown in

Figure 4-3 and Figure 4-4, and the passive LC-equivalent shown in Figure 4-2, and Fig.

4 have the same end-to-end transfer function, i.e., to response. The internal nodes

are scaled for the purpose of maximizing the dynamic range or minimizing the noise

and hence

are scaled versions of respectively.

Figure 4-2. LC ladder prototype and passive component values.

Figure 4-3. Single-ended circuit for gm-C topology.

C3dVin

V1 V3 V5

C5d

C4d

L4d

C2d

L2dRS

RL

K

I2a

I2b

I2a

I2b

K = 2 RS = 500 ohm RL = 500 ohm

L2d = 83.24 μH L4d = 60.53 μH

C1d

C1d = 481.52 pF C3d = 622.51 pF C5d = 383.15 pF

C2d = 76.42 pF C4d = 221.92 pF

K = 2 RS = 500 ohm RL = 500 ohm

L2d = 2.08 μH L4d = 2.77 μH

C1d = 12.771 pF C3d = 20.75 pF C5d = 16.05 pF

C2d = 7.3972 pF C4d = 2.548 pF

Active-RC fc = 1 MHz gm-C fc = 30 MHz

C6

C1

C7

gmK

gm0

gm2 gm4

gm1 gm3

C2

gm6 gm8

gm5 gm7

C4C3 C5

gm9

V3'

Vin V1' V2' V4' V5'

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

51

Figure 4-4. Single-ended circuit for active-RC topology.

4.3.2 Systematic Signal-Flow-Graph Based Modeling

Most active-filter design handbooks (like [30]) provide design equations that

relate the variables in the active-filter representation (shown in Figure 4-3 and Figure

4-4) to the component values in the passive LC equivalent (shown in Figure 4-2). In our

case, correct choice of independent variables is necessary to ensure that the expressions

for total noise, area and power consumption turn out to be posynomial.

We describe how the independent variables were chosen through the signal-

flow-graph (SFG) manipulations shown in Figure 4-6 to Figure 4-9. The first step in

relating the equivalent circuits involves drawing the SFG of the passive circuit. This is

shown in Figure 4-6. Next, as shown in [79], we can use the substitution and the i-shift

theorem from linear network theory [80] to get the modified signal flow graph as shown

in Figure 4-7. The i-shift theorem states that a current source between two nodes can be

equivalently represented as two current sources and two impedances, one each between

each node and ground. After this, we introduce four new design variables: and .

and have units of resistance, while and have units of transconductance. These

gain elements are introduced at various places in the SFG of Figure 4-7 to set the gains

between the internal nodes. This step gets us to Figure 4-8.

-1

-1

-1

-1

RK

R0

R2

R4

R6

R8

R9

R1

R3

R5

R7

C1 C2

C3

C4

C5Vin

-V1'V2' -V3'

V4' -V5'

C6

C7 C8

C9

V5'V1'

V3'

V3'

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

52

Fig

ure

4-5

. D

iffe

ren

tia

l cir

cuit

s fo

r (a

) g

m-C

to

po

log

y,

(b)

act

ive-R

C t

op

olo

gy

+V

in/2

-Vin

/2

C1

C1

C2

C2

C3

C3

C4

C4

C5

C5

R1

R1

R2

R2

R4

R4

R3

R3

R5

R5

R6

R6

R8

R8

R7

R7

R0

R0

R9

R9

RK

RK

-V1'

/2+

V1'

/2-V

3'/2

+V

3'/2

-V5'

/2+

V5'

/2

+V

2'/2

-V2'

/2+

V4'

/2-V

4'/2

C7

C7

C8

C8

C9

C9

C6

C6

+V

3'/2

-V3'

/2-V

3'/2

+V

3'/2

-V1'

/2

+V

1'/2

-V5'

/2

+V

5'/2

(b)

(a)

gm

Kg

m0

gm

1

gm

2g

m4

gm

3g

m5

gm

6g

m8

gm

7

gm

9

C1/2

C2/2

C3/2

C4/2

C5/2

C6

C6

C7

C7

+V

5'/2

-V5'

/2

+V

4'/2

-V4'

/2-V

3'/2

+V

3'/2

+V

2'/2

-V2'

/2-V

1'/2

+V

1'/2

+V

in/2

-Vin

/2

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

53

Note that these scaling gains are independent of each other and by introducing

them into the SFG, the loop gains of each of the loops remain unchanged, while only

the gains between the internal nodes is changed. The node voltages and the element

values change accordingly and the new values are as annotated.

Finally, noting that we can selectively also scale each of the three current

summing nodes, we introduce three additional scaling variables: and . This

gets us to Figure 4-9, and completes the signal-flow-graph manipulations. Note that in

all the SFGs shown in Figure 4-6 to Figure 4-9, all quantities within the red-outline

boxes represent a transconductance, while all quantities within the green-outline boxes

represent scalars. Also, solid lines represent voltages, and dotted lines represent currents.

Figure 4-6. Signal-flow-graph of circuit in Figure 4-2.

Figure 4-7. Modified SFG of circuit in Figure 4-2.

K

Rs

-1

1

Rs

+ sC1d

1

sL2d

+ sC2d

sL4d

+ sC4d

1-1

sC3d

-1

1

RL

+ sC5d

-V1 V3 -V5

-(I2a + I2b)Vin

I4a + I4b

K

Rs

-1

1

Rs

+s(C1d+C2d)

1

sL2d sL4d

1-1

s(C2d+C3d+C4d)

-1

1

RL

+s(C4d+C5d)

-V1 V3 -V5

-I2bVinI4b

V3 V3-V1-V5

sC2d sC2dsC4dsC4d

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

54

Figure 4-8. Modified SFG with internal node-voltage scaling.

Figure 4-9. Final, fully-scaled SFG of circuit in Figure 4-2.

The voltage and current summing nodes in the SFG shown in Figure 4-9 can

now be replaced by their equivalent gm-cell and op-amp based circuit counter-parts

using the transformations shown in Figure 4-10.

-1

1

Rs

+s(C1d+C2d)

1

1

1

1

sL2d sL4d

-1

s(C2d+C3d+C4d)

-1

1

RL

+s(C4d+C5d)

-V1 -V5

-I2b

Vin

I4b

-V5

C2ds(C2dAB)

C4ds(C4dCD)

K/Rs

A

A

B

C

DC

ABCD

-V1

ABCD

V3

CD

V3

CD

V3

CD

BCD D

AB CD

BsL2d DsL4d

s s

ABCD

-1

1

Rs/g1

+ sg1(C1d+C2d)

g1

1

g3

1

sL2d sL4d

-1

sg3(C2d+C3d+C4d)

-1

1

RL/g5

+ sg5(C4d+C5d)

-V1 -V5

-I2b

Vin

I4b

-V5

g1C2d

s(g3C2dAB)g3C4d

s(g5C4dCD)

A

A

g3B

C

g5DC

ABCD

-V1

ABCD

V3

CD

V3

CD

V3

CD

BCD D

AB CD

BsL2d DsL4d

-V1'= V3'= -V5'=

-V2’= V4’=

V3'=-V1'=-V5'=V3'=

g1K/Rs

ABCD

s s

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

55

Figure 4-10. Relating the SFG to circuit elements for gm-C and active-RC topologies.

If we now write the node voltages of the gm-C and the active-RC filter shown in

Figure 4-3 and Figure 4-4 as:

and

, then using the transformation shown in Figure 4-10, we will get the following

relationships:

gm-C topology: We find that the scaling variables and are not independent

of each other: while So if we write , then we

can take the five scaling factors and the capacitors and as the

independently selectable design variables (seven total). After that, we can express

the rest of the design variables (the eleven transconductances

and five capacitors ) in terms of these seven design variables and the

ten constants from the passive LC equivalent circuit ( -

The complete design equations are then given by (the independent variables are

henceforth written in bold in all the equations):

(4-4)

Vx

Vo

-Vy

X YVx

Vo

-11

Ro+ sCo

-Vo

Vr

sCr

Vp -VqQP

C

-Vy-1

-Vq

-Vo

Co

Ro

Q

1

sYC

1

sXC

VpP

Vr

Cr

XsC YsCVx Vy

Vo

C

P QVp VqVo

Co - Cr

Vr

Cr

Ro

1

gm gmV gmV V gmVNote:

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

56

active-RC topology: In this case, we find that the scaling variables and are

independent of each other. Consequently, we can take the seven scaling factors

and the capacitors and as the independently selectable

design variables (nine in all). After that, we can express the rest of the design

variables (the eleven resistors and seven capacitors

) in terms of these nine design variables and the ten constants

from the passive LC equivalent circuit ( - The complete

design equations are then given by:

(4-5)

It should be noted that the equations (4-4) and (4-5) are general and are not

limited to our particular elliptic transfer function or cut-off frequency. We can select

any 5th

order low-pass filter response (elliptic, Butterworth, Chebychev, Bessel, etc.),

any cut-off frequency, any ripple specification and any source and load resistance

values, and then calculate the component values for the corresponding LC equivalent

circuit. After that, the relationships in (4-4) can be used to construct the gm-C filter

circuit or the relationships in (4-5) can be used to construct the active-RC filter. Note

that for Butterworth, Chebychev or Bessel filter responses, there are no capacitors in the

series branch, i.e., and . We can also extend the analysis presented in

Figure 4-6 to Figure 4-8 for a different filter order and create a corresponding set of

equations as in (4-4) and (4-5).

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

57

4.4 Design and Optimization of the gm-C Filter Topology

The area, power consumed and total integrated noise are functions of the circuit

parameters and thus of the seven independent variables . The

following calculations are based on the complete differential circuit shown in Figure

4-5 (a).

4.4.1 Transconductor Structure

Each transconductor in a gm-C filter is linearized to improve the dynamic range

of the filter [81], [67], [68]. A typical transconductor stage is shown in Figure 4-11.

Figure 4-11. Transconductor stage.

For the transconductor, the effective transconductance is a fraction of

the actual transconductance of the output branch transistors because of the source

degeneration. Let this scaling factor be (typically of the order of 0.1) i.e.,

.

gmi

gmiV/2

gmiV/2

+Vin/2

-Vin/2

-Vout/2

+Vout/2+Vin/2 -Vin/2

-Vout/2 +Vout/2

gmai gmai

IB

VB

VDD

RS RS

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

58

4.4.2 Equations and Constraints Formation

A. Power Consumption

Let’s assume that the ratio of transconductance to the bias current of the output

branch transistors is (typically of the order of 10 V-1 . The current in each output

branch will be equal to . The total output stage

current of the transconductor will be equal to . Further, let's

assume that for each transconductor, the total current of the bias branches is a fraction

(typically around 0.3) of the current in the transconductor stage and that the power

supply voltage is . Then, the total power consumption of the transconductor is

equal to where and are constants.

Combining all the constants together as a new constant , we can write:

(4-6)

Note that has units of V2. The total power consumed in the circuit of Figure

4-5 (a) is equal to . From (4-4) and (4-6),

(4-7)

We see from (4-1) and (4-7) that the total power is a posynomial function of the

independent variables.

B. Total Area

Neglecting the area consumed by the active circuit, we assume (for simplicity)

that only the capacitors contribute to the total area. We assume that the capacitance

density on chip, i.e., the capacitance per unit area, is a constant, = fF/μm2 (a

typical value used in [71]). From the circuit in Figure 4-5 (a) and the equations in (4-4),

the total area of the filter is

(4-8)

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

59

We see from (4-1) and (4-8) that the total area is also a posynomial function of the

independent variables.

C. Total Integrated Noise

The total integrated noise of the differential circuit of Figure 4-5 (a) is twice the

total integrated noise of the single ended version of Figure 4-3. This is because the

circuit is symmetric and is sensed differentially [7].

The only noise sources in the gm-C filter are the channel thermal noise of the

transistors that make up the transconductor and the thermal noise of any degeneration

resistor that are used for linearization of the transconductor. As shown in [67], the noise

from the transconductor can be expressed as a single current source at the output of

the transconductor with a single-sided noise current spectral density given by:

(4-9)

where is the effective transconductance and is the excess noise factor which,

given a particular transconductor architecture, is just a scaling factor used to express the

total noise of the transconductor in terms of its effective transconductance. (In [67], this

factor is taken to be 2.3.)

Coming back to Figure 4-3, we see that the current noise sources for

and come in parallel and thus, these noise sources see the same current to

output voltage transfer function. Similarly, we see that and and

and , and form sets of noise sources that come in parallel and we can

write the total integrated noise of the filter at the output as

(4-10)

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

60

where are the current to voltage transfer functions from the five nodes

of the circuit to the output node, respectively. Nodal analysis on the circuit of Figure

4-3 shows that

(4-11)

where the coefficients of the polynomials are all functions only of

the constants . If we now write for ,

(4-12)

then, because the coefficients of all the polynomials for the above five integrals are

constants, we can use the closed-form expressions of Table 2-1 to calculate these five

numbers. For our particular filter shown in Figure 4-2, with the value for the constants

- as shown in the figure, these constants are computed to be

Ω2 Hz Ω

2/Hz

Ω2/Hz (4-13)

From (4-11) and (4-13), the noise integrals are

(4-14)

From (4-4), (4-10) and (4-14), the total integrated noise of the differential filter is given

by:

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

61

(4-15)

We now see from (4-1) and (4-15) that the total integrated noise is also a posynomial

function of the independent variables.

4.4.3 GP for Optimization of gm-C Filter Design

A. Dynamic Range Maximization

In the design of the 5th

order elliptic filter, the variables represent the

inter-stage gain “knobs” that we can use to vary the gains to the internal nodes. Say we

choose kΩ and mS. Figure 4-12 shows the plot of the gains (in

dB) from the input to the circuit nodes i.e., ,

, ,

and

.

Figure 4-12. Gain to internal nodes in the gm-C topology before node-voltage scaling.

We see that the gain from the input to the output node (node 5) is as expected: 0

dB pass-band gain with 0.5 dB ripple and cut-off frequency at 30 MHz. However, we

also see that the maximum gains of the internal nodes 1, 2, 3 and 4 are larger than 0 dB.

The output swings of the transconductors are limited because of headroom issues. This

means that as the input voltage is increased, because some of the internal nodes swing

more than the output, these internal transconductors will saturate earlier than the other

transconductors. This effectively reduces the dynamic range of the filter. A solution to

15 20 25 30 35-10

-5

0

5

10

15

20

25

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

62

this problem is to use the knobs and scale the gains such that all the internal

nodes swing to the same maximum amount. This is known as node voltage scaling [70],

the procedure for which is described below:

Let us define 4 new quantities: . We go back to the

passive LC equivalent in Figure 4-2, and for a given input voltage, find the absolute

maximum value of the voltages and the currents , in the pass-band. The

four new constants are related to these maximum values as follows:

(4-16)

Note that for a given passive filter design, with a constant set of component

values - , the above four quantities will come out to be

constant. For our particular elliptic filter, Ω mS ,

Ω, mS. If we now design the filter with

, the node voltages are properly scaled and we get a plot

like the one shown in Figure 4-13.

Figure 4-13. Gain to internal nodes in the gm-C topology after node voltage scaling.

This setting ensures that all transconductors swing the same amount and use the

entire voltage headroom. To then maximize the dynamic range, we must minimize the

integrated noise at the output.

15 20 25 30 35-10

-5

0

5

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

63

Because we have now fixed the four variables , we see from (4-7),

(4-8) and (4-15), that the area, power consumed and integrated noise are now

posynomial functions only of three independent variables and the constants

- and . We can thus

write the area, power and noise functions as , and

. We can then construct GPs and optimize the filter for minimum

noise under area and power constraints, or minimum area under noise and power

constraints or minimum power under noise and area constraints.

As an illustration, let's say that the area is constrained to and the power

is constrained to . Then, the geometric program to minimize the integrated

noise can be written as:

minimize:

subject to:

(4-17)

This program can be globally solved using yalmip to yield the minimum noise

design. As an example, say that we take fF μm2

S/A V and , and with the rest of the constants defined as earlier.

Further, let's assume that we want the power consumed to be less than mW

and the area to be less than μm2. Solving the above geometric

program yields a minimum mean squared output noise voltage of V2 with

pF, pF and Also, at this optimum point, the total

area is equal to but the power consumption is further reduced from its limit of

mW to mW, a saving. This implies that the design is area limited and not

power limited.

The GP in (4-17) can be solved for different and settings leading

to the Pareto-optimal curve shown in Figure 4-14.

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

64

Figure 4-14. Pareto-optimal curve for the 5th

order elliptic gm-C filter.

B. Total Integrated Noise Minimization

Going back to the plots in Figure 4-12 and Figure 4-13, we see that the inter-

stage gains are scaled so that the peak gain from the input to all the internal nodes is 0

dB. Such a scaling increases the gain from the major noise sources to the output [67]

increasing the output referred noise. However, if we are ready to accept a particular

amount of gain peaking in a partially scaled filter, the total integrated noise at the output

of the filter can be lowered by gaining the first stages up and then gaining the later

stages down. Because the end-to-end gain of the filter remains unchanged, the input

referred noise also gets lowered. The gain peaking that is introduced because of partial

scaling can be a problem when strong interferers come in. But, in an RF front-end

where such a filter will be used, there is usually some pre-filter gain that is controlled

by an automatic gain control (AGC) loop. The AGC can detect the strong interferers

and lower the signal as desired [67], thus maintaining a linear signal at the output.

Going back to the nodal analysis for the elliptic gm-C filter, we see that the peak

gains to the internal nodes are given by:

5 50 5000

20

40

60

80

100

120

Powermax

[mW]

Nm

in [

x1

0-8

V2]

Areamax

= 5000 (m)2

Areamax

= 2500 (m)2

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

65

(4-18)

where the constants are given by (4-16). Note that for the node

voltage scaled filter where , all the peak

gains in (4-18) become equal to unity, as expected.

Next, say we would like to optimize the filter for the lowest noise given a

particular maximum peak gain to the internal nodes, say , so that

(4-19)

The constraint in (4-19) is the same as the set of constraints below

(4-20)

From (4-18) and (4-20), the optimization problem can be written as a GP shown

in (4-21). Note that, compared to the GP in (4-17), we have four additional constraints

and four additional optimization variables.

minimize:

subject to:

(4-21)

Note that with , the above GP will yield the same design as the GP in (4-17). The

above GP can again be globally solved using yalmip to yield the minimum noise design.

As an illustration example, say that we take (i.e., it is acceptable to have

dB of gain peaking in the filter). The rest of the constants are the same

as defined earlier. Further, let's assume that as in the previous sub-section, we want the

power consumed to be less than mW and the area to be less than μm2. Solving

the above GP yields a minimum integrated squared noise voltage to be V2,

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

66

with the optimized pF, pF, and the optimized gains

Ω, mS, Ω, mS. Fig. 13 shows the plot

of the voltage gains from the input to the internal nodes for this design. As expected, the

first few nodes are gained up and the last node is attenuated to improve the total

integrated noise.

Also, at this optimum point, the total area is equal to but the power

consumption is reduced from its limit of mW to mW, a saving. We see

that for the same power consumption and area, the noise is reduced from

V2 to V

2, an 8.1 dB reduction in the noise when we chose

However, note that now the maximum internal node gain is 10 dB, implying that the

output swings 10 dB less. Consequently, the dynamic range reduces by 1.9 dB.

Figure 4-15. Gain to internal nodes in the gm-C topology for

This is further illustrated in the plot of the relative changes in the total integrated

noise and the dynamic range of the filter, as shown in Figure 4-16.

15 20 25 30 35-10

-5

0

5

10

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

67

Figure 4-16. Relative change in total integrated noise and dynamic range in the gm-C topology as a

function of .

4.4.4 MIGP for Dynamic Range Optimization

Suppose we wish to set mW in the GP in (4-17). Solving the GP,

we find that the area required is µm2 and the total integrated noise is

V2. We also get a set of final transconductor values as follows:

mS, mS, mS, mS,

mS, mS, mS, mS, mS,

mS, mS. It is practically very difficult to implement each

transconductor to have these different values, which are not allowed to vary much (e.g.,

due to mismatch) to preserve the filter’s transfer function. In order to mitigate this

problem, most designs use the same transconductance in each cell [67], [68] and scale

the capacitors accordingly.

However, such a design approach comes with a severe drop in dynamic range,

as illustrated next. Assuming that the maximum differential, peak-to-peak voltage swing

at the output of each transconductor is limited to 2 V, the dynamic range in the case

discussed above will be 72.99 dB. Instead of this case, we construct another GP where

we enforce all the transconductors to be the same, say (except, , to

ensure and solve the GP as shown in (4-22) below:

minimize:

subject to:

0 2 4 6 8 10-8

-6

-4

-2

0

2

M [dB]

Rel

ativ

e C

han

ge

[dB

]

Total Integrated Noise

Dynamic Range

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

68

(4-22)

Solving the above GP leads to a design with mS. While the total area

remains nearly the same, the total integrated noise improves to V2. However,

if we look at the gains to the internal nodes, shown in Figure 4-17, we find that the

maximum gain to an internal node is now 11.75 dB instead of 0 dB. Consequently, the

dynamic range of this design is only 66.76 dB. This implies a drop of 6.23 dB in the

dynamic range for the same area and power budget, because of the design choice of

making all the transconductors equal.

Figure 4-17. Gains to the internal nodes in the gm-C topology, all transconductors equal.

While the outcome of such a design choice is not beneficial for the dynamic

range of the filter, it may still be desired in a robust practical solution. Therefore, it is

worth investigating if there are ways to improve the dynamic range. Specifically, a

better outcome is possible by allowing that all transconductors can be integral multiples

of a unit transconductor. We fix the maximum integral multiple to a practically possible

value (called , and then make the integer multiples (called and the value

of the unit transconductor’s the design variable. We write:

(4-23)

15 20 25 30 35-10

-5

0

5

10

15

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

69

We select the eleven ( , and as the new design variables.

Using (4-4) and (4-23) we can express the old design variables and to

the new design variables as follows:

(4-24)

Using the relationships in (4-24), the expressions for power consumption, total

area, and total integrated noise given in (4-7), (4-8), and (4-15) respectively, can now

be written in terms of the new design variables.

Using (4-4) and (4-23), we also find that not all of these new design variables

are independent of each other, and are related by the equations

(4-25)

Finally, an MIGP can be constructed as shown in below:

minimize:

subject to:

and for (4-26)

Solving the above MIGP for , mW, we find that the

area required is µm2 and the total integrated noise is V

2. If

we look at the gains to the internal nodes, as shown in Figure 4-18, we find that the

maximum gain to any internal node is 4.3 dB, and consequently, the dynamic range of

this design has increased to 71.5 dB. The improvement in dynamic range as a function

of is plotted in Figure 4-19. We thus find that by using an MIGP formulation

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

70

and optimizing the set of transconductors to be integral multiples, the dynamic range

can be improved for the same power and area budget. In our particular design problem,

by using , the dynamic range performance can be brought to within 1.5

dB of the case where continuous transconductance values are allowed.

Figure 4-18. Gains to internal nodes in the gm-C topology when all transconductors are integral

multiples of a unit transconductor,

Figure 4-19. Dynamic range as a function of in the gm-C topology.

15 20 25 30 35-10

-5

0

5

10

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

1 3 5 7 9 11 13 15 1765

67

69

71

73

75

Scalemax

Max

imu

m D

ynam

ic R

ange

[d

B]

All gm

s equal, Dyn. Range = 66.76 dB

Continuous Scaling, Dyn. Range = 72.99 dB

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

71

4.5 Design and Optimization of active-RC Filter Topology

The area, power and total integrated noise in the active-RC topology are

functions of the circuit parameters and thus of the nine independent variables

. The following calculations are based on the circuit shown

in Figure 4-5 (b). Note that some of the design equations and techniques, and the

simulated results discussed in this section will be similar to the ones discussed above for

the gm-C filter, and will be discussed only briefly in order to minimize repetition.

4.5.1 Op-Amp Structure

All the five op-amps used in the active-RC topology shown in Figure 4-5 (b) are

assumed to be replicas of each other. Each op-amp is assumed to be implemented as a

Miller-compensated two-stage amplifier [26] as shown below in Figure 4-20.

Figure 4-20. Op-amp structure.

4.5.2 Equations and Constraints Formation

and represent the first-stage and second-stage transconductance

respectively, and are considered to be design variables that need to be optimized. The

op-amp’s finite gain and bandwidth lead to a change in the filter pass-band ripple and

the cut-off frequency. This causes the filter response to deviate from the ideal response.

For an active-RC topology implemented using nested loops as shown in Figure 4-4, the

-3dB frequency for the op-amp is usually kept slightly beyond the filter cut-off

Cm RmRm

Vip Vim

V1m

V1p

Vop Vom

Cm

Vip

Vim

Vom

Vop

gm1 gm1

gm2 gm2

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

72

frequency . This ensures that the op-amps provide a constant, high open-loop gain in

the filter pass-band, thereby limiting any deviation in the closed-loop pass-band ripple

characteristics of the filter. Given a specification of MHz in the elliptic filter, the

op-amp’s -3dB frequency is set to 2 MHz.

The distribution of the op-amp’s open-loop gain between its two stages can itself

be made a part of the optimization routine. However, we make the simplifying

assumption that each stage in the op-amp provides a DC gain of 15, giving a total op-

amp DC gain of 225. To the first order, the op-amp’s -3dB cut-off frequency can

be written as

(4-27)

From (4-27), we can write the expression for the Miller capacitance to be

(4-28)

The value of is thus set by the value of obtained through optimization.

sets the op-amp noise contribution as discussed ahead in this section. It can be

decreased only to a point where the op-amp noise starts to limit the overall noise

performance of the filter. is constrained by the op-amp second-stage DC gain and

the total amount of resistance each op-amp needs to drive. If all the capacitors

in the filter are scaled up to improve the noise performance, the corresponding resistors

need to be scaled down. Then, in order for the op-amp to be able to drive these reduced

load resistances, needs to be scaled up, increasing the power consumption.

Because we have assumed that the second-stage gain needs to be at least 15, we can

write an inequality constraint for each of the five op-amp outputs as shown in (4-29)

below.

(4-29)

A. Power Consumption

Assuming a value of 18 S/A for both stages in the op-amp, and assuming

that the current in the bias branches and CMFB stages is a fraction (typically around

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

73

0.3) of the total current in the last stage, and that the power supply voltage is , the

total power consumption of the differential circuit shown in Figure 4-5 (b) can be

written as,

(4-30)

The power consumption is thus a posynomial function of and .

B. Total Area

In case of the active-RC circuit, the chip-area is assumed to be dominated

mostly by the five Miller capacitors, the eleven resistors , and the nine

filter capacitors . From (4-5) and (4-28), we see that the total area is a function

of and .

We assume that the capacitance density on chip is a constant, . We take this

number equal to fF/μm2 as in the gm-C case. Similarly, we assume that the resistance

density on chip (Rden) is equal to 1 k /μm2. Then,

(4-31)

We see that the total area is also a posynomial function of the design variables.

C. Total Integrated Noise

There are two kinds of noise sources in the active-RC filter - the thermal noise

from the resistors and the thermal noise from the op-amps.

For computing the resistor noise, it is plausible to assume that the op-amps are

ideal, i.e. the op-amp gain and bandwidth are infinite. Resistor noise can be represented

by a noise current source, with a current PSD of , in parallel with the

corresponding resistor. Looking at the filter schematic shown in Figure 4-4, we see that

all the resistors and thus the corresponding current noise sources are connected between

high-impedance nodes (op-amp inputs) and low-impedance nodes (op-amp outputs).

The currents can thus be taken together at the high-impedance nodes, i.e. noise sources

connected to the same op-amp input can be taken in parallel, and see the same noise

transfer function to the output. We can write,

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

74

(4-32)

where are the current to voltage transfer functions from the five

nodes of the circuit to the output node, respectively. Nodal analysis shows that:

(4-33)

where the coefficients of the polynomials are all functions only of

the constants . Similar to the gm-C topology, we can write

for :

(4-34)

Because the coefficients of all the polynomials for the above five integrals are constants,

we can use the closed-form expressions of Table 2-1 to calculate these five numbers.

For our particular filter shown in Figure 4-2, with the value for the constants

- as shown in the figure for the active-RC case, these

constants are computed to be:

Ω2 Hz Ω

2/Hz

Ω2/Hz (4-35)

Using equations (4-32) to (4-35), the total integrated noise from the resistors,

, can be written as a posynomial function of .

The thermal noise from the op-amps can be represented as equivalent voltage

noise sources connected to the input of the op-amps. The PSD of this voltage noise

source is given by , where is the excess noise factor. However, if we

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

75

again assume ideal op-amp behavior (infinite gain, infinite bandwidth), we find that the

voltage noise transfer function from the input nodes of the op-amps to the output of the

filter, does not tend to zero as the frequency tends to infinity. This implies that the total

integrated noise is infinity. Even if we do a complete analysis assuming a finite gain and

a finite bandwidth in the op-amp, the noise transfer function is still a constant until the

op-amp’s unity-gain-bandwidth (which in our case is 450 MHz), implying a very large

noise integration bandwidth.

However, in any practical design, the noise from the op-amps will be bandwidth

limited because of the blocks following the filter in the signal chain. In order to capture

this behavior, we assume [82] that the output of the filter is further filtered by a single

pole filter with a cut-off frequency at ten times the filter bandwidth. The cut-off

frequency of this assumed single-pole filter is high enough to not affect the pass-band

characteristics and the stop-band attenuation of the filter, and is low enough to faithfully

represent the filtering that the op-amp noise goes through in the signal chain succeeding

the filter.

With the single-pole filter connected to the output of the filter, if we now

perform a symbolic analysis and compute the voltage-noise transfer function from the

op-amp inputs to the output of the single-pole filter, we find that:

The noise transfer functions are now 6th

order functions,

The numerator coefficients of all five op-amp voltage noise transfer functions are

now functions of the design variables . We can thus use

the closed-form expressions of Table 2-1 to calculate the noise integrals

symbolically.

The symbolic integration shows that the total integrated op-amp noise,

can be written as a posynomial function of the design variables. Op-amp noise can

thus be treated as an objective or constraint in a GP based optimization program,

thus enabling the consideration of op-amp noise in an optimization based design

methodology for active-RC circuits.

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

76

4.5.3 Geometric Program for Noise Minimization

Similar to the gm-C case, we rely on node-voltage scaling to optimize the signal

swings. The four constants to can be computed from the passive filter

circuit. For our particular filter shown in Figure 4-2, with the value for the constants

- as shown in the figure for the active-RC case, these

constants are computed to be:

(4-36)

Finally, similar to the gm-C case, we can now construct a GP as shown below in

(4-37) for the active-RC case.

minimize:

subject to:

(4-37)

Running the above GP with mW, and µm2,

(which denotes the node-voltage scaled case for dynamic range maximization)

returns an optimal design that consumes mW, occupies an area of µm2, and

has a mean squared output noise of V2, of which the resistors contribute

V2, while the rest comes from the op-amps. All the op-amp outputs have

the same peak gain from the input as shown in Figure 4-21 below.

Similar to the gm-C case, by selecting a non-unity value of , we can trade the

output signal swing for a lower mean squared output noise voltage. Figure 4-22 shows a

plot of the relative change in the total integrated noise and dynamic range as a function

of for the active-RC topology.

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

77

Figure 4-21. Gains to the internal nodes in the active-RC topology.

Figure 4-22. Relative change in total integrated noise and dynamic range as a function of in the

active-RC topology.

Figure 4-23 shows how the noise contributions from the resistors and the op-

amps change, as the value of is varied between µW and mW, for a

fixed µm2. As more power becomes available for the op-amps, the

op-amp transconductances and scale up, making it possible for the the

resistors to scale down. Consequently, in order to maintain the time-constants, the

0.2 0.4 0.6 0.8 1 1.2 1.4-10

-8

-6

-4

-2

0

2

4

6

8

Frequency [MHz]

Inte

rnal

No

de

Gai

n [

dB

]

1 2 3 4 5

0 2 4 6 8 10-8

-6

-4

-2

0

2

M [dB]

Rel

ativ

e C

han

ge

[dB

]

Total Integrated Noise

Dynamic Range

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

78

capacitors scale up, thereby improving the noise performance of the filter. However,

because the relative area consumption of the capacitors is higher than that of the

resistors, the total area increases steadily as the power consumption increases. Around

mW, the design becomes area limited and consequently, the mean

squared output noise voltage and the power consumption of the filter saturates.

In an active-RC filter implemented in a CMOS process, the on-chip resistors can

vary by more than % due to temperature or process variations. If all the resistors

vary by the same amount, while the shape of the frequency response remains the same,

the cut-off frequency of the filter shifts. If the resistors are not implemented by the same

unit cell, each resistor might vary by a different amount, and such a mismatch causes

the shape of the frequency response to change. In order to limit this change and improve

the design, we can constrain the resistors to be integral multiples of a unit resistor, and

formulate the GP as an MIGP. While the equations for and results of running this MIGP

are not provided here, it can be set up in a manner similar to the MIGP described in

Section 4.4.4 for the gm-C filter, and a robust, practical active-RC filter can be designed.

Figure 4-23. Change in noise contributions from the resistors and from the op-amps in the active-

RC topology as a function of

4.6 Conclusion

Continuous-time analog-filters are among the most important blocks in

electronic systems. In this chapter, we presented a general design and optimization

10-2

10-1

100

101

102

10-1

100

101

102

103

Powermax

[mW]

No

ise

min

[x1

0-8

V2 ]

Total Noise Resistor Noise Amplifier Noise

10-2

10-1

100

101

102

10-1

100

101

102

103

Powermax

[mW]

No

ise

min

[x1

0-8

V2 ]

Total Noise Resistor Noise Amplifier Noise

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

79

methodology that helps us find the optimal gm-C or active-RC implementation for a

given filter specification. The limitations of previously known heuristics-based and

mathematics-based design approaches are overcome by the methodology presented in

this chapter. The right choice of design variables helps us express the multiple, inter-

related performance metrics of the filter as posynomial functions, leading to the

formulation of the optimization problem as a GP that can be quickly solved to find the

globally optimal solution. We also described how the GP can be extended to an MIGP

that further improves the robustness of the filters, by enforcing the variation-sensitive

components like transconductors and resistors to be integral multiples of unit

components.

Chapter 4 - Design and Optimization of Continuous-Time, gm-C and Active-RC Filters

80

81

CHAPTER 5

Design and Optimization of a Gain-Tunable

TIA for a MEMS Oscillator

5.1 Background

A timing reference is an essential part of almost all electronic systems. Shown in

Figure 5-1 is the block diagram of a representative system-on-chip (SoC) which can be

a microprocessor, or a wireline or wireless communication chip.

Figure 5-1. Timing reference in a system-on-chip.

Such SoCs consist of a phase locked loop that is comprised of a high-frequency,

tunable, on-chip oscillator (implemented using an LC or ring configuration), locked to a

lower frequency, very stable reference oscillator. This reference oscillator comprises of

on-chip sustaining circuits connected to an off-chip resonator as shown in Figure 5-1.

When a resonator is struck by an impulse, it outputs an electrical signal that is a

decaying sinusoid at a fixed frequency. This period is set either by the properties of the

material with which the resonator is made, or by the geometrical construction of the

N

fc

fc/N = fo

Resonator

Quartz Crystal

BAW, SAW

Si MEMS

fo

Other Blocks

System-on-Chip

PDCPLF

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

82

resonator, or both. The sustaining circuits interface with the resonator to provide a fixed

frequency signal indefinitely.

For a long time, these resonators have been implemented using quartz crystals.

Quartz resonators exhibit excellent temperature stability, but are bulky and costly. As an

alternative, we can use microelectromechanical system (MEMS) based resonators such

as the surface acoustic wave (SAW) resonator [83], the bulk acoustic wave (BAW)

resonator [84], or silicon-based MEMS resonators [85]. While the SAW and BAW

resonators require a piezoelectric material for transduction of the mechanical energy

into electrical signal, silicon-based resonators are capacitively-transduced and can be

fabricated in the same process in which silicon CMOS ICs are fabricated. In addition,

silicon-based resonators are very small, reducing the cost of system-in-package based

modules. As an illustration, Figure 5-2 shows the micrograph of a Wi-Fi module that

comprises of a 5mm 5.2mm WLAN SoC and a 2mm 1.8mm quartz crystal

resonator. If, instead of using the quartz resonator, a silicon MEMS resonator is used, it

will occupy a much smaller area (0.5mm 0.5mm), and can potentially be

implemented on chip. In this chapter, we deal with such a silicon-based capacitively-

transduced MEMS resonator and describe the design and optimization of a CMOS TIA

based sustaining circuit that interfaces with this resonator to create a reference oscillator

[86].

Figure 5-2. WLAN system-in-package module.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

83

5.2 Breathe-Mode Ring Resonator

The resonator used in our design, shown in Figure 5-3, is driven differentially

between the and terminals and the output current is sensed differentially

between the and terminals. The moveable ring sections, shown in green,

vibrate in unison, going out and in together, between the fixed and

terminals.

Figure 5-3. Breathe-mode ring resonator.

The resonator is fabricated in the single-crystal silicon device layer of an SOI

wafer, and vacuum packaged using an epitaxial silicon thin film encapsulation process

[87]. This significantly reduces air damping and the quality factor is expected to be

>100,000. The dimensions of the ring resonator are tabulated below:

Transduction gap 1.5 µm

Beam length 202 µm

Beam width 12 µm

Inner radius 59 µm

Device thickness 20 µm

Table 5-1. Ring resonator dimensions.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

84

Near the resonance frequency, a lumped electrical model can be used to

represent the resonator. A single-ended version of this model is shown in Figure 5-4.

represents the feedforward capacitance between the drive and sense terminals. The

differential structure of the ring resonator reduces this parasitic feedforward capacitance

considerably [88], and it is neglected in the analysis ahead. denotes the pad

capacitance of the drive and sense terminals. and respectively represent the

motional inductance of the resonator and are related to the resonance frequency, as

follows:

(5-1)

The series resistance of the resonator is represented by the motional resistance

. It is beneficial for this motional resistance to be small. For the same drive voltage

strength, a resonator with a smaller will generate a larger sense current.

Piezoelectrically-transduced resonators like the SAW and BAW resonators, typically

have very small (around 50 to 500 ) compared to the capacitively-transduced

silicon resonators, that have orders of magnitude larger (5 to 50 ). In the case

of capacitively-transduced resonators, we can show that [89] depends on the quality

factor , the resonator bias voltage , and the transduction gap , as follows:

(5-2)

We see that has a strong dependence on . With 1.5 µm in our case, is

expected to be very high. As will be shown in Section 5.5, with V, we

measured a resonant frequency of 19.8386 MHz, 160,000 and 65 kΩ.

Figure 5-4. Electrical model for resonator near resonance.

CMRMLM

CP CP

CF

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

85

5.3 TIA Based Series Resonant Oscillators

By connecting a transimpedance amplifier (TIA) that converts the sense current

into a drive voltage and connecting it in series with the resonator in positive feedback, a

series resonant oscillator can be constructed as shown in Figure 5-5 below. So long as

the forward gain of the TIA overcomes the series motional resistance through the

resonator, there will be oscillations at the resonant frequency.

Figure 5-5. Series-resonant oscillator.

The TIA has a finite bandwidth, which is kept higher than the oscillation

frequency. This implies that at the oscillation frequency, there is a finite phase shift

through the TIA, which is to be balanced by an equal and opposite phase shift through

the resonator. This finite phase shift is kept small (less than by using the

conventional rule-of-thumb that the TIA bandwidth needs to be kept more than ten-

times the frequency of oscillation [89]. If we go by this design rule, a TIA for the

breathe mode ring resonator with an of 65 at = 20 MHz, will require a TIA

with a gain of greater than 65 and a bandwidth of at least 200 MHz. This implies a

very high gain-bandwidth product and consequently, a lot of power will dissipated in

the TIA to achieve this specification.

In order to find a practical approach to designing a low-power TIA for MEMS

oscillators, we performed a symbolic analysis of series resonant oscillators to quantify

how the TIA phase shift affects the oscillation frequency, the required TIA gain and the

phase noise of the oscillator.

CMRMLM

Vdrive

TIA

drivesense

isense

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

86

5.3.1 Symbolic Analysis of Oscillation Frequency, Required TIA Gain

and Phase Noise

Neglecting , the single-ended model of the oscillator can be drawn as shown

in Figure 5-6. The pad capacitors are lumped into the TIA and its gain can

be written as . Also, it is assumed that at the oscillation frequency , the input

and output resistance of the TIA are small compared to . This means that the TIA

does not degrade the of the resonator [90].

Figure 5-6. Series-resonant oscillator model.

At , the total phase in the loop is zero. Assuming that the TIA phase shift is

so small that ,

(5-3)

where . Also, at , the loop gain should be unity. It can be shown

that this requires

(5-4)

The achievable in capacitively-transduced micro-mechanical resonators can

be greater than 100,000. From (5-3), we see that even for a conservative of 50,000

changes by only 6 ppm for a relatively large phase shift . From (5-4), we also

CMRMLM

CP CP

isense Vdrive

RTIAejφ

isense

4kBT/RM

in2/Δf

drivesense

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

87

see that at this phase shift, needs to be only 15% more than for sustaining

oscillations.

Having looked at how the TIA phase shift affects the oscillation frequency and

the gain requirement, we next look at how it affects the phase noise requirement. In

typical electrical oscillators such as a ring oscillator or LC cross-coupled oscillator,

active devices periodically turn off or saturate, and the injected noise is cyclostationary.

Linear time variant (LTV) analysis is then required to evaluate the phase noise [91]. In

MEMS resonator based oscillators, in order to prevent severe nonlinear effects induced

by large drive amplitudes [92], an automatic level control (ALC) loop is typically used

to limit the swing to values smaller than the supply rails. The active devices in the TIA

are always on, and the injected noise is stationary. In such a case, a simple linear time

invariant (LTI) analysis [93] can be used.

In the case of the oscillator shown in Figure 5-6, the two sources of noise are the

thermal noise from the resonator, which can be represented as a current noise source

with a power spectral density (PSD) equal to in parallel with [92], and

the input referred current noise of the TIA with a PSD equal to . The phase noise

at an offset from the oscillation frequency can then be evaluated using

the LTI analysis. Under the simplifying assumptions that and ,

the expression for phase noise is

(5-5)

For small phase shifts, such that and , the phase noise plot

can be drawn as shown in Figure 5-7. The close to carrier phase noise at low-offsets is

shaped by the resonator and hence will be very low, while the phase noise floor does

not depend on Q. From (5-5), we see that this floor is directly proportional to , and is

independent of the quality factor. For a fixed , the floor will degrade for higher

. Coming back to the effect of , we see that the phase noise floor is simply scaled

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

88

by . This means that for a phase shift of , the phase noise floor will

degrade by 33% (only 1.24 dB).

Figure 5-7. Phase noise in series-resonant oscillators.

In summary, what the phase noise analysis implies is that a larger phase shift

does not cause as large a degradation of phase noise as the TIA’s noise itself. It is thus

more beneficial to invest power in reducing the TIA noise, than to increase the

bandwidth in order to limit the phase shift.

5.3.2 TIA Bandwidth

The TIA in a MEMS oscillator is typically implemented as a two-stage amplifier

consisting of a high-gain core stage followed by a unity gain buffer, usually a source

follower, to buffer the large drive terminal’s pad capacitance. The buffer stage

contributes about to of phase shift. Having established that a phase shift of

about is enough in a practical design, we investigated how the remaining

phase shift through the core TIA can be limited to under , without investing a lot of

power.

Because the power of a TIA scales with its bandwidth, the bandwidth versus

phase shift relationship needs to be studied. Traditionally, a maximally flat second order

transfer function, with the pole pair’s quality factor set at , has been used for

implementing the TIA [89], [92]. As shown in Figure 5-8, with the bandwidth of the

core TIA , set to about eight to ten times the oscillation frequency, the phase shift is

. However, if we superimpose the transfer function of the resonator on the same

Δω

Ph

as

e N

ois

e

TIA Noise

Resonator Noise

Total Noise

in2

ΔfRM

21

kBTRM1

2Vdrive2 Q

2

ωo2

Δω2

in2

Δf

RM21

4Q2

ωo2

Δω2

2Vdrive2

2Vdrive2

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

89

plot (denoted by the extremely narrow bandpass transfer function centered around ),

we see that the flat region between and is wasted. This is because, a large TIA gain

and low phase shift is required only in an extremely narrowband region around .

Instead of using a maximally flat transfer function, we can deliberately introduce some

peaking in the second order transfer function, by increasing the pole pair’s quality factor

to . As shown in Figure 5-8, the core TIA phase shift can then be limited to

with a reduced bandwidth that is only four times the oscillation frequency.

Figure 5-8. Maximally-flat versus peaking transfer function in the core TIA.

5.3.3 Symbolic Analysis Summary

The results of a theoretical symbolic analysis of the series resonant oscillator

performed in the previous sub-section, are summarized in Table 5-2 below. Based on

this analysis, we can pose the design of the TIA as an optimization program as shown in

Figure 5-9 below.

fcfc

-90

0

TIA

Gain

Phase

Shift

fo

φ

fo

QTIA = 1/ 2 QTIA = 2

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

90

Rules of thumb Symbolic analysis

Phase shift Phase shift is OK

Maximally flat TF Peaking in TF is good

TIA bandwidth TIA bandwidth can be smaller

Higher Q implies better phase

noise

TIA noise determines phase noise

floor

Table 5-2. Symbolic analysis summary for series resonant oscillators.

find: A low-noise TIA topology with tunable complex conjugate pole pair

minimize: TIA noise

subject to: Power 5 mW, Core phase shift 10°, TIA gain 65 kΩ

Figure 5-9. Design and optimization strategy for TIA.

5.4 TIA Design and Optimization

5.4.1 TIA Topology

The TIA topology most commonly used for MEMS applications consists of an

OTA and a feedback resistor to set the gain, as shown in Figure 5-10 [94].

Figure 5-10. Resistive-feedback TIA topology.

An appropriately sized feedback capacitor is added to achieve stability. The

gain of this TIA is given by . In order to achieve a gain of greater than 65 kΩ

from this topology, we would need an on-chip resistance of 65 kΩ, which will be

difficult to implement. In addition, the large value would necessitate a very large

gain-bandwidth specification for the OTA.

RF

Vout

isig

A0,ω0

CF

CP

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

91

Another issue is noise. The input referred current noise including the

contributions of and the OTA’s input-referred voltage noise, , is,

(5-6)

If this TIA is used in series with a MEMS resonator with a motional impedance

of , in order to construct an oscillator, the TIA gain needs to be set equal to ,

i.e. . In that case, as shown in Figure 5-7, the phase noise floor will be

proportional to the output referred noise voltage PSD:

(5-7)

Instead of the resistive feedback topology, consider the topology first introduced

in [95]. The single-ended topology for this TIA is shown in Figure 5-11 below. The

first stage consists of a high-gain current amplifier (with of gain of feeding

into a resistor . The second stage is a source follower that is used to buffer the

node so that it can drive the large pad capacitance . To first order, the gain of this

TIA is . The TIA augments the gain of the resistor by the

current gain , thereby allowing for the realization of a large TIA gain using

a smaller on-chip resistor.

Figure 5-11. TIA with tunable complex conjugate pole pair (biasing circuits not shown).

From [94], the input referred current noise PSD of the TIA shown in Figure

5-11 is given by,

Vdrive

isenseAOTA, ωOTA

CP

C1

C2

CP

RD

gm

gmS

Vout

iout

M1

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

92

(5-8)

where represents the current noise PSD of the bias current transistor connected

to the source of . The pole at the source of is placed more than ten times

higher than the frequency of oscillation, so the current noise of can be neglected

similar to the noise of a cascode device [26]. (5-8) shows that the noise from the

resistor and the bias current noise are attenuated by when referred to

the input.

If this TIA topology is now used in series with a MEMS resonator with a

motional impedance of , in order to construct an oscillator, the TIA gain needs

to be set equal to , i.e. . In that case, the output referred

noise voltage PSD that effectively determines the phase noise floor is given by

(5-9)

Comparing to (5-9) to (5-7), we see that because is about ten times smaller

than , and typically, the noise performance of the topology shown in Figure

5-11 is better than that of the resistive feedback TIA topology.

The complete transfer function, shown in (5-10) below, consists of a second-

order transfer function from the first stage, two non-dominant poles at and

, and a high frequency zero that can be neglected.

(5-10)

The two non-dominant poles appear outside the internal loop, and as long as the

value of is less than the drain-source resistance of the transistor , the TIA’s

stability is not determined by the value of [94]. Because forms only a part of the

TIA gain, the value of can be kept low, preventing any instability in the TIA.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

93

5.4.2 Design and Optimization

The phase shifts provided by the first stage and the two non-dominant poles

need to be minimized. Fixing the two poles such that their phase shift contribution is

small, say -15°, we see that for a low overall phase shift and high gain, the first stage

current amplifier with a second-order transfer function must be designed carefully. The

phase shift through this first stage is given by,

(5-11)

For a target phase shift of -10°, a maximally flat transfer function with

will require 8.1. But if we introduce slight peaking by designing for

(say) , we require only 4.1. This slightly reduces the phase margin of

the internal feedback loop, but as seen from (5-10), the bandwidth requirement of the

OTA and the first stage, , reduces and as a result, the power

consumption, decreases by four times.

The design of the TIA proceeds as follows: the source follower pole is kept at

ten times , fixing and the source follower power. Based on a target TIA input

current noise specification, the input-referred voltage noise of the OTA is fixed and this

fixes the gain, bandwidth and power of the OTA. Closed-form symbolic expressions are

derived for the phase shift, the phase margin of the internal loop, the power

consumption and the gain of the TIA. A nonlinear constrained optimization routine then

finds optimum values of the design variables , , , , , and that minimizes

the phase shift, while putting constraints on the main branch power, the phase margin of

the internal loop and the total gain of the TIA. Table 5-3 lists the final design

parameters.

4 pF 0.45 pF 65 kΩ

8 mS 3.6 pF < 27°

2.3 mS 8 kΩ 6.9 mW

Table 5-3. TIA design parameters.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

94

5.4.3 Transistor Level Implementation

The TIA described in the previous section was implemented as shown in Figure

5-12, while the OTA is implemented as a single-stage, cascoded, common-source

amplifier as shown in Figure 5-13. For all circuits, 2.5 V. All branch currents,

transistor , and types of transistors / passives are annotated in the schematics. The

bias voltages and are generated on chip using current mirrors. All sets of

transistors annotated by the same transistor name are scaled from the same unit

transistor. The capacitor is implemented using NMOS variable capacitors making the

TIA gain tunable. can be tuned from 0.7 pF to 3.9 pF, and correspondingly, the TIA

gain can be varied between 10 kΩ and 70 kΩ.

The input common mode voltage is set to 3 using an on-chip diode-

connected PMOS stack-up and the resistors as shown in Figure 5-14 (a) below. The

current noise from this bias resistor adds directly to the input-referred current noise

of the TIA. Referring back to (5-8), in order to reduce the noise contribution of the bias

resistor needs to be kept much higher than 600 kΩ. The largest

that can be implemented depends on the available area on-chip. As a compromise

between the area occupied and the noise contributed by the bias resistors , a value of

2 MΩ was chosen and each bias resistor was implemented using seventy, 2 µm wide, 45

µm long segments of N-doped poly resistors in series. The alternative, low-area

arrangement shown in Figure 5-14 (b), inspired by the design in [95] and implemented

in [94], was not chosen because of the problems induced by mismatch between and

variation of the triode resistances of the transistors.

The load is implemented using on-chip resistors as shown in Figure 5-12. A

small amount of tunability is provided by the PMOS variable resistor which is

connected in parallel with the resistor . In the non-ALC mode, the gate of is

connected to . In the ALC mode, the gate voltage of is generated by the ALC

as shown in Figure 5-15.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

95

Fig

ure

5-1

2.

Cir

cu

it s

chem

ati

c o

f th

e T

IA.

Vb

n1

Vb

n1

Vb

n2

Vb

n2

VA

LC

Vd

riv

e+

Vd

riv

e-

Vo

ut-

Vo

ut+

C2

C2

Vtu

ne

i se

ns

e+

i se

ns

e-

C1

C1

RD

RD

MR

D

i ou

t+i o

ut-

Vb

p1

Vb

p1

VC

MF

B

12

8 μ

A1

28

μAM

1M

1

64

μA

64

μA

64

μA

64

μA M

2M

2

M3

M3

M3

M3

64

0 μ

A6

40

μA

M4

M5

M5

M4

Vx

-V

x+

Tra

nsis

tor

gm/I

D

M1

M2

M3

M4

M5

18

15

13

.5

12

.5

15

Po

ly-P

oly

Ca

p.

0.4

5 p

F, 4

00

μm

2C

1

NM

OS

Ca

p, 0.7

to

3.9

pF

, 4

0 μ

m / 2

0 μ

mC

2

N+

Im

pu

rity

Re

sis

tor,

8 kΩ

, 2.4

μm

/ 1

2 μ

mR

D

PM

OS

Va

risto

r, 0

.5 to

>

80

, 1

2 μ

m / 0

.35

μm

MR

D

Vb

p2

C3

C3

R3

R3

Vd

riv

e-

Vd

riv

e+

Vb

p2

Vb

n1M

3M

3

M6

M6

M8

M8

M6

M7

M7

M7

M2

40

μA

40

μA

40

μA

16

μA

16

μA

Po

ly-P

oly

Ca

p,

0.3

pF

, 2

70

μm

2C

3

N-d

op

ed

Po

ly R

esis

tor,

10

0 kΩ

, 1

.5 μ

m / 1

20

μm

R3

M6

M7

M8

15

15

18

VD

D1 3

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

96

Fig

ure

5-1

3.

Cir

cu

it s

chem

ati

c o

f th

e O

TA

.

Tra

nsis

tor

gm/I

D

M1

0

M1

1

M1

2

M1

3

M1

4

20 8 7 15

10

M1

5

12

.5

15

M1

1M

11

M9

M9

M1

2

16

0 μ

A1

60

μA

i se

ns

e+

i se

ns

e-

Vx

-V

x+

M1

0M

10

Vb

n3

M1

3

Vb

p3

24

0 μ

A8

0 μ

A

C4

C4

R4

R4

Vx

+V

x-

Vb

n3

M1

3M

13

M1

4M

14

M1

4

M1

5M

15

M1

2M

12

M1

2

M1

1

Vb

n3VD

D2 3

VD

D2 3

40

μA

40

μA

40

μA

32

μA

16

μA

16

μA

Po

ly-P

oly

Ca

p,

0.1

pF

, 9

2 μ

m2

C4

N

-do

pe

d P

oly

Re

sis

tor,

10

0 kΩ

, 1

.5 μ

m / 1

20

μm

R4

M9

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

97

The ALC consists of pseudo-differential envelope detectors followed by a

comparator that compares the amplitude with a reference value [92]. Unlike the

single-ended implementation in [92], the noise from the ALC affects both the drive

nodes and equally, and thus the ALC does not affect the phase noise of

the differential signal.

A micrograph of the CMOS chip, wire-bonded to the MEMS die is shown in

Figure 5-16. The total active CMOS area is 0.15 mm2.

Figure 5-14. Generating the sense terminal bias.

Figure 5-15. ALC block diagram.

VDD2

3

VDD1

3 isense-

isense+

N-doped Poly Resistor, 2 MΩ,

2 μm / 45 μm, 70 series segmentsR5

R5

R5

M16

M16

M16

PMOS, 36 μm / 2 μmM16

isense-

isense+

(a) (b)

Mtriode

Mtriode

Vdrive+

625 nA

Vdrive-

625 nA5 pF

Vref

100 kΩ

10 pF

VALC

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

98

Figure 5-16. System micrograph.

5.5 Measurement Results

The gain of the TIA is first measured by driving 100 kΩ on-board resistors

connected to its inputs as shown in Figure 5-17. After adjusting for the resistor and the

gain of the on-board 50 Ω interface driver and buffer circuits, the gain of the TIA is

measured to be tunable between 12 kΩ and 69 kΩ.

Figure 5-17. Measurement of TIA gain.

Another open-loop measurement is done with the resonator connected to the

TIA. As shown in Figure 5-18, with the TIA gain set at 65 kΩ, at V, the

open-loop gain at the peak is slightly over 0 dB, implying that kΩ. It also

shows that and 19.8386 MHz.

TIAALCV

dri

ve

+

i se

ns

e+

i se

ns

e-

Vd

riv

e-

isense+isense-

100 kΩ

TIA Buffer

CMOS Chip

Vector

Network

Analyzer

Differential to

Single-Ended

Single-Ended

to Differential

Vdrive+

Vdrive-

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

99

Figure 5-18. Open-loop measurement of the MEMS resonator characteristics.

Next, the loop is closed and the signal amplitude and phase noise at the

oscillator's output are measured. Figure 5-19 shows the phase noise of the oscillator

output voltage, with and without the ALC loop engaged, on the same graph, showing

that the ALC loop does not raise the noise floor. The differential peak-to-peak swing in

both cases is 1640 mV. The swing can be controlled from 160 mV to 1700 mV by the

ALC loop.

Figure 5-19. Closed-loop measurements.

The total power consumption is 6.9 mW from a 2.5 V supply. The breakdown of

the power consumption for the various blocks is given in below.

Power consumption

First stage 1.03 mW

Second stage 2.9 mW

ALC 1.33 mW

OTA and CMFB 1.3 mW

Total power consumption 6.9 mW

Table 5-4. TIA power-consumption breakdown.

isense+isense-

TIA Buffer

CMOS Chip

Vector

Network

Analyzer

Differential to

Single-Ended

Single-Ended

to Differential

Vdrive+

Vdrive-

CMRMLM

MEMS Chip

isense+isense-

TIA Buffer

CMOS Chip

Oscilloscope /

PN Meter /

Spectrum

Analyzer

Differential to

Single-Ended

Single-Ended

to Differential

Vdrive+

Vdrive-

CMRMLM

MEMS Chip

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

100

5.6 Performance Summary and Comparison

Having talked about the performance of the MEMS oscillator presented in this

work, we next explain how we can compare our performance to other state-of-the-art

MEMS oscillators. Different MEMS oscillators have different oscillation frequencies

and motional resistances. While a figure of merit exists for comparing the performance

of high-frequency oscillators like the LC or ring [96], there is no figure-of-merit that

exists in the literature that compares the performance of reference oscillators. In the

next sub-section, we describe our attempt to devise a suitable figure of merit to compare

the circuit design effort in series-resonant oscillators with different and resonator .

5.6.1 Circuit Design Figure-of-Merit in Series-Resonant Oscillators

In a PLL, where a high frequency oscillator is locked to the reference oscillator,

the close-to-carrier phase noise of the high frequency oscillation is determined by the

thermal phase noise floor of the reference oscillator [97]. Furthermore,

because the Q is typically very high (> 1,000) in these reference oscillators, the

and regions exist only at very low offsets, and as a result, the phase noise floor

dominates. In the reference oscillator, the phase noise floor is thus the primary

performance metric.

Next, taking into consideration the DC power consumption, the phase noise

floor should be normalized to . Also, as shown in Figure 5-20, in a PLL, the

phase noise floor of the reference oscillator is scaled by the square of the divide ratio

[97]. This implies that if we use two reference oscillators with different in a PLL for

achieving the same high-frequency oscillation, for a fair comparison, the reference

oscillators' phase noise floor should be scaled by .

Finally, (5-5) shows that the noise floor scales as . But, is a specification

from the MEMS design, and is not under the circuit designer's control. Circuit design

effort is spent in designing a good TIA around it. Thus, for comparing the circuit design

between two reference oscillators with different resonator 's, the phase noise floor

should be scaled by .

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

101

Figure 5-20. Phase noise in a PLL.

In summary, a suitable figure of merit for the circuit design effort in series-

resonant oscillators can thus be written as

(5-12)

Compared to the commonly used LC oscillator figure of merit [96], the FoM in

(5-12) includes the additional effect of the resonator loss that the oscillator overcomes,

embodied here in the term . In LC oscillators, the equivalent term would be the loss

resistance in the LC resonator, which is almost entirely contributed by the series loss

resistance in the inductors. If we look closely at the LC oscillators implemented in

standard CMOS technologies, we see that even for different oscillation frequencies, the

loss resistance term does not vary by more than an octave. This is primarily because of

the fixed resistivities of the metals used to implement the inductor turns, and the limited

range of inductor values that are possible to be implemented on chip. Consequently,

when comparing different LC oscillators, neglecting the corresponding resonator loss

resistance does not severely impact the evaluation of the oscillator’s circuit design effort.

On the other hand, different series-resonant oscillators at different frequencies,

made with different kinds of MEMS resonators, exhibit widely different loss resistances,

. (5-2) shows that depends on multiple factors like the transduction gap, the

resonator bias voltage, the tensile strength of the material used to construct the resonator,

Phase

Detector

+

Loop

Filter

N

fo

fc

fc/N = fo

£(fo)

£(f)

£(f)

N2 * £(fo)

Frequency

Divider

Reference

Oscillator

Tunable High-

Frequency Oscillator

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

102

etc. Different MEMS resonators can thus have different , usually spread across 4-5

orders of magnitude. Consequently, it becomes imperative to include the impact of

when comparing the design effort of different series-resonant oscillators.

5.6.2 Performance Comparison

This power-frequency-loss-normalized phase-noise FoM shown in (5-12) has

units of and is used in Table 5-5 to compare the oscillator in this work with

previously published designs. The MEMS oscillator presented in this work achieves the

highest FoM among the capacitively-transduced MEMS oscillators and also competes

well with the piezoelectrically-transduced MEMS oscillator designed in [98]. Note that

while the oscillator in [92] achieves similar phase noise at thrice the while burning

only a tenth of the power, the resonator loss is about forty times less compared to the

oscillator in this work. Consequently, this work achieves a better power-frequency-loss-

normalized phase-noise FoM.

Ref. No. [98] [90] [92] This work,

[86]

Transduction Piezo Capacitive

CMOS Tech. Node 0.18 µm 0.18 µm 0.35 µm 0.35 µm

[MHz] 1,006 8.29 60 20

7,100 1,040 48,000 160,000

[kΩ] 0.15 26 1.5 65

[pF] 0.5 -NA- -NA- 4

[mW] 7.2 8.7 0.78 6.9

PN floor [dBc/Hz] -154 -106 -132 -131

FoM [Hz2Ω

2] 3.27 10

19 8.74 10

14 6.77 10

17 1.36 10

19

Table 5-5. MEMS oscillator performance summary and comparison.

5.7 Conclusion

A low-noise, low-power oscillator based on a high-Q, high-loss, capacitively-

transduced MEMS resonator and a tunable, high-gain CMOS TIA has been presented.

An analytical study of the phase noise in series-resonant oscillators has been performed

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

103

and it was used to optimize the performance of the TIA. Compared to state-of-the-art

series-resonant oscillators that employ capacitively-transduced MEMS resonators, the

oscillator in this work achieves the highest circuit design figure of merit.

Chapter 5 - Design and Optimization of a Gain-Tunable TIA for a MEMS Oscillator

104

105

CHAPTER 6

Conclusions and Future Outlook

6.1 Summary

An analog circuit design problem involves finding a proper circuit solution that

optimizes multiple inter-related performance metrics like noise, gain, bandwidth, power,

area, etc. The research presented in this thesis aimed at developing fast and accurate

design and optimization methodologies for analog circuits that are capable of dealing

with complex circuits like multi-stage amplifiers and high-order filters. The goal of this

research was to demonstrate the design and optimization of different, representative

classes of analog circuits, and use them to provide a generic set of techniques. Given a

particular design problem at hand, a designer can suitably adapt these techniques to not

only find the right design, but also test rules of thumb and conventional wisdom, and

port the design to different sets of specifications and process corners.

Chapter 1 oriented the reader to the ‘one problem, multiple solution’ aspect of

analog circuit design problems. The symbolic and numerical circuit analysis approaches

were described and a case was made for the importance of using a symbolic expressions

based analog circuit design and optimization framework. Also, a brief overview of the

prior work done in the space of analog circuit and optimization was provided.

In Chapter 2, a technique to compute the total integrated noise by visual

inspection in linear, passive networks was presented. It was used to demonstrate how

the track-mode noise in a switched-capacitor stage can be computed in a small number

of steps, instead of going through a complete noise-transfer function integration. Finally,

the technique was extended to show how one can symbolically integrate a general noise

transfer function of any order to get closed-form expressions for total integrated noise.

Such expressions were not readily available and had prevented the adoption of symbolic

analysis in the design and optimization of noise limited analog circuits.

The remaining chapters described proof-of-concept examples that illustrate how

symbolic analysis can be applied to the design and optimization of representative analog

Chapter 6 - Conclusions and Future Outlook

106

blocks. Chapter 3 presented the design and optimization of a nested-Miller-

compensated three-stage operational transconductance amplifier for use in high-speed

switched-capacitor circuits. In Chapter 4, continuous-time analog-filter design and

optimization was described. A formal systematic approach to converting a given LC

passive realization to either a gm-C or an active-RC topology was provided. Closed-

form symbolic expressions were used in formulating a GP and an MIGP for the design

and optimization of these filters. The limitations of previously known heuristics-based

and mathematics-based design approaches are overcome by the methodology presented

in this chapter. Finally, Chapter 5 described the analysis, design, and measurement

results of a low-noise, low-power, series-resonant MEMS oscillator at 20 MHz.

Symbolic analysis was used to evaluate the impact of TIA bandwidth on the oscillator

frequency and phase noise, and accordingly a suitable topology is chosen and optimized.

A new figure-of-merit that faithfully captures the circuit design effort in the MEMS

oscillator was also developed.

6.2 Areas of Future Work

This work can be further extended in several different directions, as described

below.

Design and optimization tool: A software tool that can encode the techniques

described in this work would be useful for analog circuit designers. This is

especially true for the design and optimization of continuous-time analog filter

designs because of their regular structure and widespread usage. Such a tool will

need to integrate design entry and analysis through a SPICE netlist or Spectre user-

interface, to the symbolic and optimization toolbox in MATLAB. The Stanford

Circuit Optimization Tool (SCOT) [99], a tool developed to generate optimized

digital designs can be used a reference.

Computing noise by inspection in active circuits: While Chapter 2 used the Nyquist

and Bode theorem to show how we can visually compute the total integrated noise,

the method is applicable only for passive circuits because the Nyquist theorem

assumes thermal equilibrium. In order to come up with a similar method for active

circuits, we need to modify the Nyquist and Bode theorems for the non-equilibrium

Chapter 6 - Conclusions and Future Outlook

107

case. While this is a tough problem to solve, any solution that gives insight into the

computation of noise in the active circuits without the need to integrate, will be a

significant contribution.

A more holistic approach to filter design and optimization: In Chapter 4, we

described the design and optimization of gm-C and active-RC filters by starting from

an LC equivalent passive circuit, where the value of its components ( and ) were

fixed. In a more holistic approach, we can broaden the search space in order to find

a better design. This can be done by keeping the passive LC filter’s component

values as design variables themselves. This can be achieved by relaxing the

specifications like ripple, stop-band attenuation, etc. to be bounded in a particular

range instead of being fixed.

Chapter 6 - Conclusions and Future Outlook

108

109

APPENDIX A

Proof of the Bode Theorem for

Impedance/Admittance Integrals

The two-sided noise voltage PSD in (2-8) can be written as:

(A.1)

where denotes the complex conjugate of . is a rational function in .

Consequently, is a rational function of and hence, . Let the

auto-correlation of the noise voltage be . By the Wiener-Khintchine theorem,

and

form a Fourier transform pair [100]:

(A.2)

where denotes the inverse Fourier transform. Let denote the inverse Laplace

transform. Further, let:

(A.3)

From (A.1), (A.2), and (A.3):

(A.4)

Because there is at-least one noise-generating, dissipative resistive element in

the network, there is no imaginary axis pole pair in . In addition, for stability

reasons, none of the poles will be in the right half plane. Thus, can have complex

conjugate pole pairs with negative real parts, can have poles on the negative real axis,

can have a pole at , and zeros anywhere on the s-plane. In addition, the number of

zeros will be at least one less than the number of poles. Otherwise, as , the value

of will tend to a finite number or keep on increasing, and from (2-8)

and (2-10), the total integrated noise will become unbounded. can thus be written

as follows, after partial fraction expansion:

Appendix A – Proof of the Bode Theorem for Impedance / Admittance Integrals

110

(A.5)

Similarly, can be written as follows:

(A.6)

The inverse Fourier and inverse Laplace transforms of the terms in (A.5) and (A.6) are

[101]:

(A.7)

where and are respectively the signum and step functions defined as

follows:

for for

(A.8)

. (A.9)

The equations in (A.7) effectively show that,

(A.10)

Furthermore, because is causal, from (A.7), (A.8), (A.9) and (A.10):

for

for

(A.11)

for

for

(A.12)

From (A.4), (A.11), and (A.12):

Appendix A – Proof of the Bode Theorem for Impedance / Admittance Integrals

111

(A.13)

Substituting into (A.13), we get,

(A.14)

The initial value theorem [102] states that,

(A.15)

Also, from (A.6), we get,

(A.16)

From (A.14), (A.15) and (A.16):

(A.17)

Substituting into (A.2) in we get,

(A.18)

Finally, from (2-10), (A.17), and (A.18), we get:

(A.19)

For the case where has no poles at DC, i.e. , (A.19)

simplifies to the result in [47], given in (2-11).

A similar analysis leads to the corresponding results for the mean squared

current noise.

Appendix A – Proof of the Bode Theorem for Impedance / Admittance Integrals

112

113

APPENDIX B

Solution of Equation (2-30)

Let and respectively denote the largest even and odd integers

smaller than or equal to . Using (2-31), we can re-write the expressions for ,

, and as follows:

(B.1)

Using the expressions in (B.1), (2-30) can be re-written as:

(B.2)

Substituting and simplifying,

Appendix B – Solution of Equation (2-30)

114

(B.3)

At this point, we can use the following polynomial identities:

(B.4)

where it is understood that and for and for . Also,

according to the way the functions and were defined, the following holds:

(B.5)

Using (B.4) and (B.5) in (B.3):

(B.6)

Because (2-30) has to hold for all , (B.6) has to hold for all . Equating the coefficients

of the corresponding orders of on both sides of (B.6) will lead to m independent

equations to solve for the coefficients of :

,

......

......

Appendix B – Solution of Equation (2-30)

115

(B.7)

The above equations can be written in matrix form as follows:

(B.8)

where the matrices , and are defined below in (B.9). and are square

matrices of size whereas and are row matrices of size .

=

(B.9)

where for and for , and denotes the modulus

(remainder) of after division by 2.

Let us define a new column matrix such that:

=

(B.10)

From (B.8) and (B.10):

Appendix B – Solution of Equation (2-30)

116

(B.11)

Using Cramer's rule [103] on the matrix relationship (B.11), is equal to the

ratio of determinants of two matrices and :

(B.12)

where the matrix is equal to , and the matrix is formed by replacing the last

column of by the column matrix

117

Bibliography

[1] “iPhone 5 Technical Specifications.” Available:

http://www.apple.com/iphone/specs.html [Accessed: 26-Mar-2013].

[2] M. Barros, J. Guilherme, and N. Horta, Analog circuits and systems

optimization based on evolutionary computation techniques. Berlin: Springer-

Verlag, 2010.

[3] F. V. Fernandez, A. Rodrigue-Vazquez, J. L. Huertas, and G. G. Gielen,

Symbolic analysis techniques: applications to analog design automation.

Wiley-IEEE Press, 1998.

[4] “The SPICE Home Page.” Available:

http://cutler.eecs.berkeley.edu/Classes/icbook/SPICE [Accessed: 12-Mar-

2013].

[5] “Virtuoso Spectre Circuit Simulator.” Available:

http://www.cadence.com/products/cic/spectre_circuit/pages/default.aspx

[Accessed: 12-Mar-2013].

[6] “Symbolic Math Toolbox.” [Online]. Available:

http://www.mathworks.com/products/symbolic/ [Accessed: 12-Mar-2013].

[7] A. Dastgheib and B. Murmann, “Calculation of total integrated noise in

analog circuits,” IEEE Transactions on Circuits and Systems I: Regular

Papers, vol. 55, no. 10, pp. 2988–2993, Nov. 2008.

[8] “BSIM Group >> BSIM4.” [Online]. Available:

http://www-device.eecs.berkeley.edu/bsim/?page=BSIM4 [Accessed: 20-

Mar-2013].

[9] F. Silveira, D. Flandre, and P. G. A. Jespers, “A gm/ID based methodology for

the design of CMOS analog circuits and its application to the synthesis of a

silicon-on-insulator micropower OTA,” IEEE Journal of Solid-State Circuits,

vol. 31, no. 9, pp. 1314–1319, Sep. 1996.

[10] “Optimization Toolbox - MATLAB.” [Online]. Available:

http://www.mathworks.com/products/optimization/ [Accessed: 12-Mar-2013].

[11] N. Damera-Venkata and B. L. Evans, “An automated framework for

multicriteria optimization of analog filter designs,” IEEE Transactions on

Circuits and Systems II: Analog and Digital Signal Processing, vol. 46, no. 8,

pp. 981–990, 1999.

Bibliography

118

[12] R. Harjani, R. A. Rutenbar, and L. R. Carley, “OASYS: a framework for

analog circuit synthesis,” IEEE Transactions on Computer-Aided Design of

Integrated Circuits and Systems, vol. 8, no. 12, pp. 1247–1266, Dec. 1989.

[13] M. G. R. Degrauwe, O. Nys, E. Dijkstra, J. Rijmenants, S. Bitz, B. L. A. G.

Goffart, E. A. Vittoz, S. Cserveny, C. Meixenberger, G. Van Der Stappen,

and H. J. Oguey, “IDAC: an interactive design tool for analog CMOS

circuits,” IEEE Journal of Solid-State Circuits, vol. 22, no. 6, pp. 1106–1116,

Jun. 1987.

[14] H. Y. Koh, C. H. Sequin, and P. R. Gray, “OPASYN: a compiler for CMOS

operational amplifiers,” IEEE Transactions on Computer-Aided Design of

Integrated Circuits and Systems, vol. 9, no. 2, pp. 113–125, Feb. 1990.

[15] W. Nye, D. C. Riley, A. Sangiovanni-Vincentelli, and A. L. Tits,

“DELIGHT.SPICE: an optimization-based system for the design of integrated

circuits,” IEEE Transactions on Computer-Aided Design of Integrated

Circuits and Systems, vol. 7, no. 4, pp. 501–519, Apr. 1988.

[16] E. S. Ochotta, R. A. Rutenbar, and L. R. Carley, “Synthesis of high-

performance analog circuits in ASTRX/OBLX,” IEEE Transactions on

Computer-Aided Design of Integrated Circuits and Systems, vol. 15, no. 3, pp.

273–294, Mar. 1996.

[17] G. Van der Plas, G. Debyser, F. Leyn, K. Lampaert, J. Vandenbussche, G. G.

E. Gielen, W. Sansen, P. Veselinovic, and D. Leenarts, “AMGIE - a synthesis

environment for CMOS analog integrated circuits,” IEEE Transactions on

Computer-Aided Design of Integrated Circuits and Systems, vol. 20, no. 9, pp.

1037–1058, Sep. 2001.

[18] M. Krasnicki, R. Phelps, R. A. Rutenbar, and L. R. Carley, “MAELSTROM:

efficient simulation-based synthesis for custom analog cells,” Proceedings of

the 36th Design Automation Conference, DAC 1999, pp. 945–950.

[19] R. Phelps, M. Krasnicki, R. A. Rutenbar, L. R. Carley, and J. R. Hellums,

“Anaconda: simulation-based synthesis of analog circuits via stochastic

pattern search,” IEEE Transactions on Computer-Aided Design of Integrated

Circuits and Systems, vol. 19, no. 6, pp. 703–717, Jun. 2000.

[20] M. J. Krasnicki, R. Phelps, J. R. Hellums, M. McClung, R. A. Rutenbar, and

L. R. Carley, “ASF: a practical simulation-based methodology for the

synthesis of custom analog circuits,” Proceedings of the IEEE/ACM

International Conference on Computer Aided Design, ICCAD 2001, pp. 350–

357.

[21] A. Doboli and R. Vemuri, “Exploration-based high-level synthesis of linear

analog systems operating at low/medium frequencies,” IEEE Transactions on

Bibliography

119

Computer-Aided Design of Integrated Circuits and Systems, vol. 22, no. 11,

pp. 1556–1568, Nov. 2003.

[22] “Virtuoso NeoCircuit - Datasheet.” [Online]. Available:

http://www.cadence.com/rl/Resources/datasheets/VirNeoCircuit_ds.pdf

[Accessed: 03-Apr-2013].

[23] M. Horowitz, “Digital Analog Design.” [Online]. Available: http://www.dr-

assoc.net/IEEE-CEDA-DAC-2012/IEEE-CEDA-DAC-2012.html [Accessed:

29-Mar-2013].

[24] D. Patil, S. Yun, S.-J. Kim, A. Cheung, M. Horowitz, and S. Boyd, “A new

method for design of robust digital circuits,” Proceedings of the 6th

International Symposium on Quality of Electronic Design, ISQED 2005, pp.

676–681.

[25] M. Hershenson, S. Boyd, and T. H. Lee, “Optimal design of a CMOS op-amp

via geometric programming,” IEEE Transactions on Computer-Aided Design

of Integrated Circuits and Systems, vol. 20, no. 1, pp. 1–21, Jan. 2001.

[26] P. R. Gray, P. J. Hurst, and S. H. Lewis, Analysis and design of analog

integrated circuits. Wiley, 2009.

[27] Y. P. Tsividis and C. Mcandrew, Operation and modeling of the MOS

transistor. Oxford University Press, Incorporated, 2011.

[28] V. Vasudevan, “A simple technique to evaluate the noise spectral density in

operational amplifier based circuits using the adjoint network theory,” IEEE

Transactions on Circuits and Systems I: Regular Papers, vol. 51, no. 11, pp.

2175–2178, Nov. 2004.

[29] P. Feldmann and R. W. Freund, “Circuit noise evaluation by Pade

approximation based model-reduction techniques,” Proceedings of the

IEEE/ACM International Conference on Computer-Aided Design, ICCAD

1997, pp. 132–138.

[30] R. Schaumann, H. Xiao, and M. E. V. Valkenburg, Design of analog filters.

Oxford University Press, 2009.

[31] G. Groenewold, “Noise and group delay in active filters,” IEEE Transactions

on Circuits and Systems I: Regular Papers, vol. 54, no. 7, pp. 1471–1480, Jul.

2007.

[32] L. Toth, G. Efthivoulidis, V. Gopinathan, and Y. P. Tsividis, “General results

for resistive noise in active RC and MOSFET-C filters,” IEEE Transactions

on Circuits and Systems II: Analog and Digital Signal Processing, vol. 42, no.

12, pp. 785–793, Dec. 1995.

Bibliography

120

[33] G. Efthivoulidis, L. Toth, and Y. P. Tsividis, “Noise in gm-C filters,” IEEE

Transactions on Circuits and Systems II: Analog and Digital Signal

Processing, vol. 45, no. 3, pp. 295–302, Mar. 1998.

[34] S. Koziel, R. Schaumann, and Haiqiao Xiao, “Analysis and optimization of

noise in continuous-time OTA-C filters,” IEEE Transactions on Circuits and

Systems I: Regular Papers, vol. 52, no. 6, pp. 1086–1094, Jun. 2005.

[35] J. H. Fischer, “Noise sources and calculation techniques for switched

capacitor filters,” IEEE Journal of Solid-State Circuits, vol. 17, no. 4, pp.

742–752, Apr 1982.

[36] R. Schreier, J. Silva, J. Steensgaard, and G. C. Temes, “Design-oriented

estimation of thermal noise in switched-capacitor circuits,” IEEE

Transactions on Circuits and Systems I: Regular Papers, vol. 52, no. 11, pp.

2358–2368, Nov. 2005.

[37] T. Kailath, A. H. Sayed, and B. Hassibi, Linear estimation. Prentice Hall,

2000.

[38] K. Kundert, The designer’s guide to SPICE and Spectre®, 1st ed. Springer,

1995.

[39] A. Jeffrey and D. Zwillinger, Table of integrals, series, and products.

Academic Press, 2007.

[40] E. Kreyszig, Advanced engineering mathematics. John Wiley & Sons, 2010.

[41] J. S. Bay, Fundamentals of linear state space systems. WCB/McGraw-Hill,

1999.

[42] R. S. Martin, G. Peters, and J. H. Wilkinson, “The QR algorithm for real

hessenberg matrices,” Numerical Mathematics, vol. 14, no. 3, pp. 219–231,

Feb. 1970.

[43] R. H. Bartels and G. W. Stewart, “Solution of the matrix equation AX+ XB=

C [F4],” Communications of the ACM, vol. 15, no. 9, pp. 820–826, Sep. 1972.

[44] M. Abdalla, R. Wang, and R. McLauchlan, “Solving Lyapunov equations

symbolically,” Proceedings of the IEEE International Symposium on

Computer-Aided Control System Design, 1996, pp. 504–509.

[45] H. Nyquist, “Thermal agitation of electric charge in conductors,” Physical

Review, vol. 32, pp. 110–113, Jul. 1928.

[46] R. Q. Twiss, “Nyquist’s and Thevenin’s theorems generalized for

nonreciprocal linear networks,” Journal of Applied Physics, vol. 26, no. 5, pp.

599–602, May 1955.

Bibliography

121

[47] A. Papoulis, Probability, random variables, and stochastic processes.

McGraw-Hill, 1984.

[48] B. Murmann, “Thermal noise in track-and-hold circuits: analysis and

simulation techniques,” IEEE Solid-State Circuits Magazine, vol. 4, no. 2, pp.

46–54, Jun. 2012.

[49] G. C. Newton, Jr., L. A. Gould, and J. F. Kaiser, Analytical design of linear

feedback controls. Wiley, 1957, App. E.

[50] S.-C. Lee, Y.-D. Jeon, J.-K. Kwon, and J. Kim, “A 10-bit 205-MS/s 1.0-mm2

90-nm CMOS pipeline ADC for flat panel display applications,” IEEE

Journal of Solid-State Circuits, vol. 42, no. 12, pp. 2688–2695, Dec. 2007.

[51] J. H. Hiujsing, “Multi-stage amplifier with capacitive nesting for frequency

compensation,” U.S. Patent 455950217, Dec. 1985.

[52] K. N. Leung, P. K. T. Mok, and W.-H. Ki, “Right-half-plane zero removal

technique for low-voltage low-power nested Miller compensation CMOS

amplifier,” Proceedings of the 6th IEEE International Conference on

Electronics, Circuits and Systems, ICECS 1999, vol. 2, pp. 599–602.

[53] R. Nguyen and B. Murmann, “The design of fast-settling three-stage

amplifiers using the open-loop damping factor as a design parameter,” IEEE

Transactions on Circuits and Systems I: Regular Papers, vol. 57, no. 6, pp.

1244–1254, Jun. 2010.

[54] A. Pugliese, F. A. Amoroso, G. Cappuccino, and G. Cocorullo, “Settling time

optimization for three-stage CMOS amplifier topologies,” IEEE Transactions

on Circuits and Systems I: Regular Papers, vol. 56, no. 12, pp. 2569–2582,

Dec. 2009.

[55] S. Seth and B. Murmann, “Settling time and noise optimization of a three-

stage operational transconductance amplifier,” Proceedings of the IEEE

International Symposium on Circuits and Systems, ISCAS 2012, pp. 205–208.

[56] S. Seth and B. Murmann, “Settling time and noise optimization of a three-

stage operational transconductance amplifier,” IEEE Transactions on Circuits

and Systems I: Regular Papers, vol. 60, no. 5, pp. 1168-1174, May 2013.

[57] T. Konishi, K. Inazu, J. G. Lee, M. Natsui, S. Masui, and B. Murmann,

“Design optimization of high-speed and low-power operational

transconductance amplifier using gm/ID lookup table methodology,” IEICE

Transactions on Electronics, vol. E94–C, no. 3, pp. 334–345, Mar. 2011.

[58] R. J. Apfel and P. R. Gray, “A fast-settling monolithic operational amplifier

using doublet compression techniques,” IEEE Journal of Solid-State Circuits,

vol. 9, no. 6, pp. 332–340, Jun. 1974.

Bibliography

122

[59] B. Y. T. Kamath, R. G. Meyer, and P. R. Gray, “Relationship between

frequency response and settling time of operational amplifiers,” Solid-State

Circuits, IEEE Journal of, vol. 9, no. 6, pp. 347–352, Jun. 1974.

[60] S. K. Mitra, Analysis and synthesis of linear active networks. J. Wiley, 1969.

[61] P. V. A. Mohan, VLSI analog filters: active RC, OTA-C, and SC. Springer,

2013.

[62] “Analog filters using MATLAB - Toolbox.” [Online]. Available:

http://www.es.isy.liu.se/publications/books/Analog_Filters_Using_MATLAB

/download/toolbox.html [Accessed: 05-Jan-2011].

[63] T. Deliyannis, Y. Sun, and J. K. Fidler, Continuous-time active filter design.

CRC Press, 2010.

[64] A. I. Zverev, Handbook of filter synthesis. Wiley-Interscience, 2005.

[65] J. Mahattanakul, P. Khumsat, and W. Surakampontorn, “Ladder-simulation

elliptic bandpass active-RC filter structure employing identical resistors,” IET

Circuits, Devices & Systems, vol. 3, no. 4, p. 187, Apr. 2009.

[66] T. Laxminidhi, V. Prasadu, and S. Pavan, “A low power 44-300 MHz

programmable active-RC filter in 0.18 μm CMOS,” Proceedings of the IEEE

Custom Integrated Circuits Conference, CICC 2007, pp. 683–686.

[67] F. Behbahani, W. Tan, A. Karimi-Sanjaani, A. Roithmeier, and A. A. Abidi,

“A broad-band tunable CMOS channel-select filter for a low-IF wireless

receiver,” IEEE Journal of Solid-State Circuits, vol. 35, no. 4, pp. 476–489,

Apr. 2000.

[68] A. J. Lewinski and J. Silva-Martinez, “A 30-MHz fifth-order elliptic low-pass

CMOS filter with 65-dB spurious-free dynamic range,” IEEE Transactions

on Circuits and Systems I: Regular Papers, vol. 54, no. 3, pp. 469–480, Mar.

2007.

[69] T. Gao, W. Li, N. Li, and J. Ren, “A 80–400 MHz 74 dB-DR gm-C low-pass

filter with a unique auto-tuning system,” Proceedings of the 16th Asia and

South Pacific Design Automation Conference, ASP-DAC 2011, pp. 115–116.

[70] K. Martin and A. S. Sedra, “Designing leap-frog and SFG filters with

optimum dynamic range,” Proceedings of the IEEE, vol. 65, no. 8, pp. 1210–

1211, Aug. 1977.

[71] C. Myers, B. Greenley, D. Thomas, and U. Moon, “Continuous-time filter

design optimized for reduced die area,” IEEE Transactions on Circuits and

Systems II: Express Briefs, vol. 51, no. 3, pp. 105–110, Mar. 2004.

Bibliography

123

[72] G. Groenewold, “Optimal filter networks,” IEEE Transactions on Circuits

and Systems II: Express Briefs, vol. 55, no. 9, pp. 843–847, Sep. 2008.

[73] G. Groenewold, “Optimal ladder filters,” IEEE Transactions on Circuits and

Systems II: Express Briefs, vol. 56, no. 2, pp. 147–151, Feb. 2009.

[74] J. F. Fernandez-Bootello, M. Delgado-Restituto, and A. Rodriguez-Vazquez,

“Matrix methods for the dynamic range optimization of continuous-time gm-C

filters,” IEEE Transactions on Circuits and Systems I: Regular Papers, vol.

55, no. 9, pp. 2525–2538, Oct. 2008.

[75] S. Boyd, S.-J. Kim, L. Vandenberghe, and A. Hassibi, “A tutorial on

geometric programming,” Optimization and Engineering, vol. 8, no. 1, pp.

67–127, Apr. 2007.

[76] E. L. Lawler and D. E. Wood, “Branch-and-bound methods: a survey,”

Operations Research, vol. 14, no. 4, pp. 699–719, Jul. 1966.

[77] J. Lofberg, “YALMIP : a toolbox for modeling and optimization in

MATLAB,” Proceedings of the IEEE International Symposium on Computer

Aided Control Systems Design, 2004, pp. 284–289.

[78] J. Lofberg, “YALMIP : a toolbox for modeling and optimization in

MATLAB.” [Online]. Available: http://users.isy.liu.se/johanl/yalmip/

[Accessed: 12-Jun-2011].

[79] M. A. Tan and R. Schaumann, “A reduction in the number of active

components used in transconductance grounded capacitor filters,”

Proceedings of the IEEE International Symposium on Circuits and Systems,

ISCAS 1990, pp. 2276–2278.

[80] L. Chua, Linear and non linear circuits. McGraw-Hill Education, 2000.

[81] E. Sánchez-Sinencio and J. Silva-Martinez, “CMOS transconductance

amplifiers, architectures and active filters: a tutorial,” IEE Proceedings of

Circuits, Devices and Systems, vol. 147, no. 1, pp. 3–12, Jan. 2000.

[82] F. Trofimenkoff, D. Treleaven, and L. Bruton, “Noise performance of RC-

active quadratic filter sections,” IEEE Transactions on Circuit Theory, vol. 20,

no. 5, pp. 524–532, May 1973.

[83] N. A. Ramli and A. N. Nordin, “Design and modeling of MEMS SAW

resonator on Lithium Niobate,” Proceedings of the 4th International

Conference on Mechatronics, ICOM 2011, pp. 1–4.

[84] G. G. Fattinger, “BAW resonator design considerations - An overview,”

Proceedings of the IEEE International Frequency Control Symposium, FCS

2008, pp. 762–767.

Bibliography

124

[85] J. C. Salvia, Micro-oven based temperature compensation systems for MEMS

oscillators. Stanford University Ph.D. Thesis, 2010.

[86] S. Seth, S. Wang, T. Kenny, and B. Murmann, “A -131 dBc/Hz, 20 MHz

MEMS oscillator with a 6.9 mW, 69 k-ohm, gain-tunable CMOS TIA,”

Proceedings of the European Solid-State Circuits Conference, ESSCIRC 2012,

pp. 249–252.

[87] S. Wang, S. A. Chandorkar, A. B. Graham, M. W. Messana, J. Salvia, and T.

W. Kenny, “Encapsulated mechanically coupled fully-differential breathe-

mode ring filters with ultra-narrow bandwidth,” Proceedings of the 16th

International Solid-State Sensors, Actuators and Microsystems Conference

(TRANSDUCERS), 2011, pp. 942–945.

[88] S. A. Bhave, D. Gao, R. Maboudian, and R. T. Howe, “Fully-differential

poly-SiC Lame mode resonator and checkerboard filter,” Proceedings of the

18th IEEE International Conference on Micro Electro Mechanical Systems,

2005, pp. 223–226.

[89] C.-C. Nguyen and R. T. Howe, “An integrated CMOS micromechanical

resonator high-Q oscillator,” IEEE Journal of Solid-State Circuits, vol. 34, no.

4, pp. 440–455, Apr. 1999.

[90] F. Nabki and M. N. El-Gamal, “A high gain-bandwidth product

transimpedance amplifier for MEMS-based oscillators,” Proceedings of the

34th European Solid-State Circuits Conference, 2008, pp. 454–457.

[91] T. H. Lee and A. Hajimiri, “Oscillator phase noise: A tutorial,” IEEE Journal

of Solid-State Circuits, vol. 35, no. 3, pp. 326–336, Mar. 2000.

[92] Yu-Wei Lin, Seungbae Lee, Sheng-Shian Li, Yuan Xie, Zeying Ren, and C.

T.-C. Nguyen, “Series-resonant VHF micromechanical resonator reference

oscillators,” IEEE Journal of Solid-State Circuits, vol. 39, no. 12, pp. 2477–

2491, Dec. 2004.

[93] J. Craninckx and M. Steyaert, “Low-noise voltage-controlled oscillators using

enhanced LC-tanks,” IEEE Transactions on Circuits and Systems II: Analog

and Digital Signal Processing, vol. 42, no. 12, pp. 794–804, Dec. 1995.

[94] J. Salvia, P. Lajevardi, M. Hekmat, and B. Murmann, “A 56Mohm CMOS

TIA for MEMS applications,” Proceedings of the IEEE Custom Integrated

Circuits Conference, CICC 2009, pp. 199–202.

[95] B. Razavi, “A 622 Mb/s 4.5 pA/√ Hz CMOS transimpedance amplifier for

optical receiver front-end,” Digest of Technical Papers, IEEE International

Solid-State Circuits Conference, ISSCC 2000, pp. 162–163.

Bibliography

125

[96] D. Ham and A. Hajimiri, “Concepts and methods in optimization of

integrated LC VCOs,” IEEE Journal of Solid-State Circuits, vol. 36, no. 6, pp.

896–909, Jun. 2001.

[97] J. Craninckx and M. Steyaert, Wireless CMOS frequency synthesizer design.

Springer, 1998.

[98] H. M. Lavasani, W. Pan, B. Harrington, R. Abdolvand, and F. Ayazi, “A 76

dBOhm 1.7 GHz 0.18 um CMOS tunable TIA using broadband current pre-

amplifier for high frequency lateral MEMS oscillators,” IEEE Journal of

Solid-State Circuits, vol. 46, no. 1, pp. 224–235, Jan. 2011.

[99] “The Stanford Circuit Optimization Tool.” [Online]. Available:

http://code.google.com/p/thescot/ [Accessed: 08-May-2013].

[100] A. B. Carlson, P. B. Crilly, and J. C. Rutledge, Communication systems, 4th

ed. McGraw-Hill, 2002.

[101] B. M. Project and U. S. O. of N. Research, Tables of integral transforms:

Based, in part, on notes left by Harry Bateman. McGraw-Hill, 1954.

[102] K. H. Lundberg, H. R. Miller, and R. L. Trumper, “Initial conditions,

generalized functions, and the laplace transform troubles at the origin,” IEEE

Control Systems, vol. 27, no. 1, pp. 22–35, Jan. 2007.

[103] R. A. Horn and C. R. Johnson, Matrix analysis. Cambridge, U.K.: Cambridge

University Press, 1985.